Point-contact Andreev-reflection spectroscopy of doped manganites: Charge carrier spin-polarization and proximity effects (Review Article)

Materials with spin-polarized charge carriers are the most demanded in the spin-electronics. Particularly requested are the so-called half-metals which have the maximum attainable value of carrier spin polarization. Doped manganites are in the list of compounds with, potentially, half-metallic pro...

Ausführliche Beschreibung

Gespeichert in:
Bibliographische Detailangaben
Datum:2013
Hauptverfasser: Krivoruchko, V.N., D’yachenko, A.I., Tarenkov, V.Yu.
Format: Artikel
Sprache:English
Veröffentlicht: Фізико-технічний інститут низьких температур ім. Б.І. Вєркіна НАН України 2013
Schriftenreihe:Физика низких температур
Schlagworte:
Online Zugang:http://dspace.nbuv.gov.ua/handle/123456789/118223
Tags: Tag hinzufügen
Keine Tags, Fügen Sie den ersten Tag hinzu!
Назва журналу:Digital Library of Periodicals of National Academy of Sciences of Ukraine
Zitieren:Point-contact Andreev-reflection spectroscopy of doped manganites: Charge carrier spin-polarization and proximity effects (Review Article) / V.N. Krivoruchko, A.I. D’yachenko, V.Yu. Tarenkov // Физика низких температур. — 2013. — Т. 39, № 3. — С. 276–292. — Бібліогр.: 118 назв. — англ.

Institution

Digital Library of Periodicals of National Academy of Sciences of Ukraine
id irk-123456789-118223
record_format dspace
spelling irk-123456789-1182232017-05-30T03:04:42Z Point-contact Andreev-reflection spectroscopy of doped manganites: Charge carrier spin-polarization and proximity effects (Review Article) Krivoruchko, V.N. D’yachenko, A.I. Tarenkov, V.Yu. К 75-летию со дня рождения И. К. Янсона Materials with spin-polarized charge carriers are the most demanded in the spin-electronics. Particularly requested are the so-called half-metals which have the maximum attainable value of carrier spin polarization. Doped manganites are in the list of compounds with, potentially, half-metallic properties. The point-contact (PC) Andreevreflection (AR) spectroscopy is a robust and direct method to measure the degree of current spin polarization. In this report, advances in PCAR spectroscopy of ferromagnetic manganites are reviewed. The experimental results obtained on “classic” s-wave superconductor — ferromagnetic manganites PCs, as well as related theoretical models applied to deduce the actual value of charge carrier spin-polarization, are discussed. Data obtained on “proximity affected” contacts is also outlined. Systematic and repeatable nature of a number of principal experimental facts detected in the AR spectrum of proximity affected contacts suggests that some new physical phenomena have been documented here. Different models of current flow through a superconductor–half-metal ferromagnet interface, as well as possibility of unconventional superconducting proximity effect, have been discussed. 2013 Article Point-contact Andreev-reflection spectroscopy of doped manganites: Charge carrier spin-polarization and proximity effects (Review Article) / V.N. Krivoruchko, A.I. D’yachenko, V.Yu. Tarenkov // Физика низких температур. — 2013. — Т. 39, № 3. — С. 276–292. — Бібліогр.: 118 назв. — англ. 0132-6414 PACS: 72.25.Mk, 74.45.+c, 72.25.Ba http://dspace.nbuv.gov.ua/handle/123456789/118223 en Физика низких температур Фізико-технічний інститут низьких температур ім. Б.І. Вєркіна НАН України
institution Digital Library of Periodicals of National Academy of Sciences of Ukraine
collection DSpace DC
language English
topic К 75-летию со дня рождения И. К. Янсона
К 75-летию со дня рождения И. К. Янсона
spellingShingle К 75-летию со дня рождения И. К. Янсона
К 75-летию со дня рождения И. К. Янсона
Krivoruchko, V.N.
D’yachenko, A.I.
Tarenkov, V.Yu.
Point-contact Andreev-reflection spectroscopy of doped manganites: Charge carrier spin-polarization and proximity effects (Review Article)
Физика низких температур
description Materials with spin-polarized charge carriers are the most demanded in the spin-electronics. Particularly requested are the so-called half-metals which have the maximum attainable value of carrier spin polarization. Doped manganites are in the list of compounds with, potentially, half-metallic properties. The point-contact (PC) Andreevreflection (AR) spectroscopy is a robust and direct method to measure the degree of current spin polarization. In this report, advances in PCAR spectroscopy of ferromagnetic manganites are reviewed. The experimental results obtained on “classic” s-wave superconductor — ferromagnetic manganites PCs, as well as related theoretical models applied to deduce the actual value of charge carrier spin-polarization, are discussed. Data obtained on “proximity affected” contacts is also outlined. Systematic and repeatable nature of a number of principal experimental facts detected in the AR spectrum of proximity affected contacts suggests that some new physical phenomena have been documented here. Different models of current flow through a superconductor–half-metal ferromagnet interface, as well as possibility of unconventional superconducting proximity effect, have been discussed.
format Article
author Krivoruchko, V.N.
D’yachenko, A.I.
Tarenkov, V.Yu.
author_facet Krivoruchko, V.N.
D’yachenko, A.I.
Tarenkov, V.Yu.
author_sort Krivoruchko, V.N.
title Point-contact Andreev-reflection spectroscopy of doped manganites: Charge carrier spin-polarization and proximity effects (Review Article)
title_short Point-contact Andreev-reflection spectroscopy of doped manganites: Charge carrier spin-polarization and proximity effects (Review Article)
title_full Point-contact Andreev-reflection spectroscopy of doped manganites: Charge carrier spin-polarization and proximity effects (Review Article)
title_fullStr Point-contact Andreev-reflection spectroscopy of doped manganites: Charge carrier spin-polarization and proximity effects (Review Article)
title_full_unstemmed Point-contact Andreev-reflection spectroscopy of doped manganites: Charge carrier spin-polarization and proximity effects (Review Article)
title_sort point-contact andreev-reflection spectroscopy of doped manganites: charge carrier spin-polarization and proximity effects (review article)
publisher Фізико-технічний інститут низьких температур ім. Б.І. Вєркіна НАН України
publishDate 2013
topic_facet К 75-летию со дня рождения И. К. Янсона
url http://dspace.nbuv.gov.ua/handle/123456789/118223
citation_txt Point-contact Andreev-reflection spectroscopy of doped manganites: Charge carrier spin-polarization and proximity effects (Review Article) / V.N. Krivoruchko, A.I. D’yachenko, V.Yu. Tarenkov // Физика низких температур. — 2013. — Т. 39, № 3. — С. 276–292. — Бібліогр.: 118 назв. — англ.
series Физика низких температур
work_keys_str_mv AT krivoruchkovn pointcontactandreevreflectionspectroscopyofdopedmanganiteschargecarrierspinpolarizationandproximityeffectsreviewarticle
AT dyachenkoai pointcontactandreevreflectionspectroscopyofdopedmanganiteschargecarrierspinpolarizationandproximityeffectsreviewarticle
AT tarenkovvyu pointcontactandreevreflectionspectroscopyofdopedmanganiteschargecarrierspinpolarizationandproximityeffectsreviewarticle
first_indexed 2025-07-08T13:35:06Z
last_indexed 2025-07-08T13:35:06Z
_version_ 1837085972342243328
fulltext © V.N. Krivoruchko, A.I. D’yachenko, and V.Yu. Tarenkov, 2013 Low Temperature Physics/Fizika Nizkikh Temperatur, 2013, v. 39, No. 3, pp. 276–292 Point-contact Andreev-reflection spectroscopy of doped manganites: Charge carrier spin-polarization and proximity effects (Review Article) V.N. Krivoruchko, A.I. D’yachenko, and V.Yu. Tarenkov A.A. Galkin Donetsk Physics and Technology Institute National Academy of Sciences of the Ukraine 72 R. Luxemburg Str., Donetsk 83114, Ukraine E-mail: krivoruc@gmail.com Received September 18, 2012 Materials with spin-polarized charge carriers are the most demanded in the spin-electronics. Particularly re- quested are the so-called half-metals which have the maximum attainable value of carrier spin polarization. Doped manganites are in the list of compounds with, potentially, half-metallic properties. The point-contact (PC) Andreev- reflection (AR) spectroscopy is a robust and direct method to measure the degree of current spin polarization. In this report, advances in PCAR spectroscopy of ferromagnetic manganites are reviewed. The experimental results ob- tained on “classic” s-wave superconductor — ferromagnetic manganites PCs, as well as related theoretical models applied to deduce the actual value of charge carrier spin-polarization, are discussed. Data obtained on “proximity af- fected” contacts is also outlined. Systematic and repeatable nature of a number of principal experimental facts de- tected in the AR spectrum of proximity affected contacts suggests that some new physical phenomena have been documented here. Different models of current flow through a superconductor–half-metal ferromagnet interface, as well as possibility of unconventional superconducting proximity effect, have been discussed. PACS: 72.25.Mk Spin transport through interfaces; 74.45.+c Proximity effects; Andreev reflection; SN and SNS junctions; 72.25.Ba Spin-polarized transport in metals. Keywords: current spin polarization, Andreev-reflection spectroscopy, ferromagnetic manganites, unconven- tional pairing. Contents 1. Introduction .......................................................................................................................................... 277 2. Manganites and PCAR technique ......................................................................................................... 278 2.1. Ferromagnetic half-metallic manganites ...................................................................................... 278 2.2. Fabrication and characterization of point contacts ....................................................................... 278 3. Theoretical basis of the AR spectroscopy of current spin polarization ................................................ 279 3.1. Ballistic-type contacts .................................................................................................................. 279 3.2. Diffusive-type contacts ................................................................................................................ 280 3.3. Contacts with spin-dependent transmission.................................................................................. 280 4. Conventional half-metallic characteristics of point contacts ................................................................ 280 5. Proximity affected contacts, “anomalous” Andreev reflection ............................................................. 282 6. Triplet transport in proximity sSC-HMF point contacts ...................................................................... 284 6.1. Triplet pairing in sSC/F heterostructures ...................................................................................... 284 6.2. Triplet supercurrent in sSC-manganites PCs ................................................................................ 287 6.3. Latent superconductivity of doped manganites ........................................................................... 287 7. Conclusions .......................................................................................................................................... 289 References ................................................................................................................................................ 289 Point-contact Andreev-reflection spectroscopy of doped manganites: Charge carrier spin-polarization Low Temperature Physics/Fizika Nizkikh Temperatur, 2013, v. 39, No. 3 277 1. Introduction Spintronics is a research field wherein two fundamental branches of physics — magnetism and electronics — are combined [1–4], and is based on the opportunity of ferro- magnetic materials to provide spin-polarized current. The performance of many spintronics devices improves dra- matically as the degree of charger carriers spin polariza- tion, PC, increases, and materials with highly spin- polarized current are of crucial importance here. Particular attention has focused on the so-called “half-metals”. Half- metals are ferromagnets having a Fermi surface in one spin band, and a gap for the opposite spin direction; i.e., half- metals have the maximum attainable value of spin polari- zation (PC = 100%). Yet, measuring PC is not an easy task. Almost all methods of measuring it necessitate removing electrons from the material under investigation. The densi- ties of quasiparticle states at the Fermi level EF must there- fore be weighted by their probability of escape, which gen- erally depends on the measuring method and is also different for electrons with up (↑) and down (↓) spin. The degree of spin polarization can be defined in sever- al different ways. Methods available to measure spin pola- rization include spin-polarized photoemission [5], and transport measurements in point contacts and tunnel junc- tions with two ferromagnetic electrodes (see, e.g., [6]), or with ferromagnetic and superconducting electrodes. The latter case splits into two spectroscopic methods: the tunneling spectroscopy, originated in Tedrow–Meservey works on probing spin polarization [7], and the point-con- tact (PC) Andreev-reflection (AR) spectroscopy based on the way of supercurrent conversion into quasiparticle cur- rent at superconductor (SC)–normal metal (N) interface invented by Andreev [8]. The conventional technique of spin-resolved photoe- mission [5] measures the spin of the electrons emitted from a region close to the surface of a ferromagnet of order of 5–10 Å, and thus is quite surface-sensitive. The Tedrow and Meservey [7] method requires the material under in- vestigation to be fabricated as part of a ferromagnet/super- conductor tunnel junction, in which the superconducting density of states is then Zeeman split by the application of a magnetic field of several Tesla. The PC method offers several apparent advantages compared to these techniques. There are no restrictions on the sample geometry and one can avoid complex fabrication steps. In addition, it has excellent energy resolution, typically about 0.1 meV, and does not necessarily require an applied magnetic field. This spectroscopic method was invented over 35 years ago by I.K. Yanson [9] as an experimental tool to investi- gate the interaction mechanisms between electrons and pho- nons in metals. The PC technique was later used to study all kinds of scattering of electrons by elementary excitation in metals [10–12]. If one of the sides of a PC is a SC, the PC’s differential conductance contains fundamental information on the excitation spectrum of quasiparticles. For this reason PC spectroscopy has become an important, sometimes unique, tool for investigation of superconductors (see, e.g., recent review [13] and references therein). The PCAR spectroscopy is a technique in which the dif- ferential conductance G = dI/dV is measured for an elec- trical PC with little or no tunneling barrier established be- tween the superconducting tip and the counter-electrode (or vice versa). The idea that AR at the interface between a ferromagnet (F) and a SC can be used for direct probing of current spin polarization was apparently suggested for the first time by de Jong and Beenakker [14]. The idea is sim- ple: the Andreev reflection can be visualized as two cur- rents of electrons with opposite spins flowing inside a normal metal towards its interface with a SC. At the inter- face (more precisely, within the coherence length from the interface) the two currents recombine creating the current of singlet Cooper pairs. In a nonmagnetic metal both spin subband currents are the same, so in the superconducting state one can observe a 100% increase in the net current over the normal state. It was suggested [14] that in a fer- romagnetic metal the total Andreev current is defined by that spin channel where the normal-state current is smaller, because the excess electrons in the other channel will not find partners to form Cooper pairs with. The success of the first experiments for measuring charge carrier’s spin polarization [15,16] with a direct su- perconductor PC to an F established AR as a useful tool for characterizing materials for spin-electronics applications. Since then, this technique has been used to measure spin polarization in a broad range of ferromagnetic materials, including the transition metal elements Fe, Ni, and Co [15,17], metallic alloys of transition metals such as permal- loy [15,18], Heusler alloys such as NiMnSb [15,19], half- metals such as CrO2 [15,19–21], doped manganites [15,19, 22–29], and other ferromagnetic metals and semiconduc- tors. The foregoing may lead to a conclusion that the PCAR technique is indeed a robust and universal way for measuring spin polarization applicable to all magnetic compounds. There are, however, a few cases when this technique can not be applied. For example, magnetite, Fe3O4, is ferrimagnet with anomalously high Curie tem- perature ~ 850 K. The band-structure calculations (see, e.g., [30]) have predicted that Fe3O4 is a half-metal with only one spin subband at the Fermi level. Experimentally, the spin-resolved photoemission on epitaxial thin films indeed indicated a spin polarization about 80% at room temperature [31] However, below a Verwey transition at about 120 K, single crystal Fe3O4 shows a nonmetallic (polaronic) type of conductivity (see, e.g., [32]), and, in fact, the PCAR method cannot be used. The objective of the report is to give a short review of experiments exploring charge carrier’s spin-polarization in half-metallic manganites by the PCAR technique. We start, Sec. 2, with brief review of physics of ferromagnetic man- V.N. Krivoruchko, A.I. D’yachenko, and V.Yu. Tarenkov 278 Low Temperature Physics/Fizika Nizkikh Temperatur, 2013, v. 39, No. 3 ganites as materials with half-metallic band structure. Then we discuss major issues concerning the fabrication and characterization of point contacts and possible related imperfections. Section 3 is devoted to the theoretical basis for the PCAR spectroscopy of current spin polarization. The models of point contacts with different character of conduction (ballistic, diffusive, etc.) are discussed. The following three sections are central. Firstly, we overview the experimental results obtained for conventional PCs of s-wave SC-manganite and estimate the degree of charge carrier’s spin polarization for doped manganites (Sec. 4). Then, in Sec. 5, we discuss the results obtained for the so- called “proximity affected” contacts for which a few un- conventional features have been detected. A possible phys- ical scenario explaining the observations is discussed in Sec. 6. We end with summary and conclusions. 2. Manganites and PCAR technique 2.1. Ferromagnetic half-metallic manganites In doped manganites, the ferromagnetic ordering has been attributed to the double-exchange interaction between the valence electronic states of Mn3+–O2––Mn4+ bonds. Within the double-exchange model [33–36] the itinerant charge carriers provide both the magnetic interaction be- tween nearest Mn3+–Mn4+ ions and the electrical conduc- tivity. Due to the short mean free path (that is typically the distance of a few lattice parameters), the charge carrier probes the magnetization on a very short length scale. As a result, for these systems a strong interplay between local magnetic order and electric properties exists. Also, there is a large onsite Hund’s energy, which for Mn3+ is about 1.5 eV. A direct consequence of Hund’s interaction is a large spin splitting of the conduction band into majority and minority subbands in the ferromagnetic phase. In op- timally doped compounds, e.g., La0.7Ca0.3MnO3 and La0.7Sr0.3MnO3, the ge↑ (Mn) holes are itinerant while gt ↓ (Mn) electrons are localized, i.e., the ground state electrons are nearly perfectly spin polarized. Note that in general, the Fermi level EF intersects both up and down bands, so that densities of states of both the ↑ and ↓ carriers are present at EF. According to the general classification [6], when the ↑ carriers are itinerant and the ↓ carriers are localized (or visa versa), we have transport half-metal, classified as a type IIIA (or IIIB) half-metal. (The suffix A or B indicates whether the conduction elec- trons are ↑ or ↓ with respect to magnetization direction.) Thus, doped manganites are type IIIA half-metallic ferro- magnets (HMFs). For further discussion, it is noteworthy that numerous experimental results, including transport studies [33–37], indicate that doped manganites are, in fact, “bad” metals even in at low temperature. Theoretically, HMFs has been justified by band struc- ture calculations [38–43]. For the first time direct experi- mental evidence of half-metallic density of states in ferro- magnetic La0.7Ca0.3MnO3 was obtained by scanning tunneling spectroscopy by Wei et al. [44]. These results were then confirmed by optical [5,45–48] and transport measurements [15,22–29,49–55]. However, it should be taken into account that different experiments yield quite different value of spin polarization (see Table 1). The rea- sons are obvious: the electron densities of state at the Fer- mi level EF must be properly weighted and weighting fac- tor depends on the physical process used in measurements. In tunneling experiments, the weighting factors are the appropriate spin-dependent tunneling matrix elements. In PC experiments depending on the character of transport involved, either ballistic or diffusive, the density of states must be weighted either by the Fermi velocity of the elec- trons or its square, respectively [56,57], and so on. 2.2. Fabrication and characterization of point contacts A PC is simply a contact between two metals, a metal and a semiconductor, or a metal and a superconductor, with a nanoscale radius. PCs may be fabricated in a num- ber of ways [12,13]. Traditionally, most often used is the so-called “needle-anvil” configuration when the sample to be studied is one electrode and the other is a metallic tip, which is gently pressed against the sample surface. Typi- cally, the tip has an ending diameter of some tens of mi- crometers and is easily deformed during the contact. This means that, as a rule, parallel contacts are formed between the sample and the tip. In general, this is not harmful to spectroscopy, unless the sample is highly inhomogeneous on a length scale compared with tip end. The needle-anvil technique offers several apparent ad- vantages compared to other techniques: it is nondestruc- tive and several measurements can be carried out on the same sample; there are no restrictions on the sample geo- metry and one can avoid complex fabrication steps, etc. Yet, there are some objective drawbacks such as poor thermal and mechanical stability of the junction. A surface modification due to uncontrolled surface oxides or other chemical reactions on the surface of both the sample and the tip are also possible and should be taken into account, however, the impact of these effects on, e.g., PC is diffi- cult to quantify. An important characteristic of a PC is its diameter d (it is assumed that the shape of a contact is a circular orifice). Depending on the value of the d, different regimes of con- duction are possible. The diameter d of a PC with resis- tance RN can be estimated using the Wexler interpolation formula [58]: 2 16 , 3N lR dd ρ ρ ≈ + π (1) where ρ is the resistivity of the crystal, and l stands for the charge carriers mean free path. The mean free path in the metals can be estimated using the well known expression for conductivity: σ = e2N(EF)D, where D = lvF/3 is the Point-contact Andreev-reflection spectroscopy of doped manganites: Charge carrier spin-polarization Low Temperature Physics/Fizika Nizkikh Temperatur, 2013, v. 39, No. 3 279 electron diffusive constant, N(EF) stands for the density of state, and vF is the electron velocity at the Fermi surface. Using for the parameter values available from literature [33–36], one can get the mean free path for carriers in doped manganites l ~ 100 Å. Then, for a contact with typi- cal resistance RN ~ 100 Ω, we find from Eq. (1) the contact diameter d ~ 100 Å. The relation l ~ d means that the con- tact with RN ~ 100 Ω borders between the diffusive (l << d) and ballistic (l >> d) types of conductivity. Correspon- dingly, contacts with lower resistance are close to the dif- fusive type of conduction and contacts with higher resis- tance are close to the ballistic type of conduction. For contacts of N metal with SC there is an additional characteristic length. The solution of the Bogoliubov-de Gennes equations near an N/SC interface yields that the AR does not occur abruptly at the interface but over a length scale of the order of the superconducting coherence lengths ξS and ξN in both directions, respectively. In gener- al, ξS and ξN characterize the so-called “proximity” effects. Namely, ξN is the length over which superconducting cor- relations spread into N metal, and ξS is the length over which superconductivity of SC can be depressed due to contact with N metal [59]. If the contact size is considera- bly smaller than ξS proximity effects can be neglected; if not, proximity effects should be taken into account [13]. 3. Theoretical basis of the AR spectroscopy of current spin polarization As was discovered in the early 1960s, at energies below the superconducting gap, a charge transport through a normal nonmagnetic metal being in contact with a SC is possible only due to a specific process called Andreev ref- lection [8]: a two-particle process in which, in the N metal, an incident electron above the Fermi energy EF and an electron below EF with an opposite momentum and spin are coupled together and transferred across the interface in the SC side forming a Cooper pair in the condensate. Si- multaneously, an evanescent hole appears in the N metal. The charge doubling at the interface enhances the subgap conductance and this phenomenon has indeed been ob- served in the case of a perfectly transparent interface. The picture is significantly modified when spin comes into play. For a conventional s-wave pairing, a Cooper pair is spin-less (S = 0) and spin is not carried through the N/SC interface. Therefore, if the N metal is an F and there is an imbalance between spin-up and spin-down populations, the AR is suppressed and the subgap conductance can be re- duced below the normal-state value down to zero as the polarization approaches 100%. Theoretically, the behavior of normal metal–constric- tion–superconductor (N–c–SC) systems or N–SC point contacts was analyzed by Zaitsev [60]. This work was ex- panded on by Blonder et al. [61] who added the possibility of a delta-function scattering potential at the N/SC inter- face. This provides a simple means for including the effect of interfacial scattering, allowing the successful modeling of N–SC point contacts where the transport ranges from the high current density, purely ballistic or Sharvin [62] regime, to the low current density, a tunneling regime. Subsequently de Jong and Beenakker [14] pointed out that for an F–SC contact the spin polarization of the conduction electrons in the F would affect AR, because not every incident electron from the F is able to be reflected as a hole to form a Cooper pair that can move into the SC. They argued that (for an ideal F–SC contact) this will reduce the AR transmission probability to a factor (1 − PC) where PC is the polarization of the current in the F. Given the widespread interest to un- derstanding the spin-dependent current and related spintron- ic effects in heterogeneous structures, this led to the devel- opment of the PCAR method for determining the degree of current spin polarization in ferromagnetic systems of spin- tronics interest, and to the development of several different models to interpret the F–SC PC data. Below we very briefly discuss the conventional models of conductivity for a point SC–F contact: ballistic, diffu- sive, and a contact with possible spin discrimination of transmission at SC/F interface. (For more details see origi- nal reports [14,27,56,57,61,63,64], textbook [12] and re- cent review [13].) 3.1. Ballistic-type contacts The most simplified but theoretically most self- consistent is the 1D (i.e., all the involved momenta are normal to the interface) model of a PC with ballistic type of conduction. That is, the charge carriers elastic mean free path l is much larger than the diameter of the contact d and there is no scattering in the contact area. The barrier is represented by a repulsive potential located at the interface with a dimensionless amplitude, ~ Zδ(x) (here it is sup- posed that the interface lies in the yz-lane); the value Z = 0 describes a clean interface, and values of Z > 5 correspond to the tunneling type of conductivity. Note that, if the Fer- mi velocities of N and SC metals in a contact are different (vFN ≠ vFS), then the effective parameter Z ≠ 0 even if there is no scattering of carriers at the interface. It can be as- sumed that almost all the voltage V is applied directly to the contact itself and, therefore, the metals on both sides of the contact are in equilibrium state and can be described by the Fermi distribution functions. For the F–SC contact with the ballistic type of conduc- tivity, the contact conductance G can be conveniently represented as the sum of two terms [15]: ( ) (1 – ) ( ) ( ),C NS C PSG V P G V P G V= + (2) the nonpolarized part GNS (which corresponds to an An- dreev contact with nonmagnetic metal) and the completely polarized part GPS (which corresponds to a SC–ferro- magnetic metal contact with full polarization of charge V.N. Krivoruchko, A.I. D’yachenko, and V.Yu. Tarenkov 280 Low Temperature Physics/Fizika Nizkikh Temperatur, 2013, v. 39, No. 3 carriers). The Andreev conductance of the N/SC contact, GNS, is given by the well-known expression (the BTK model [61]): ( )NSG V = [ ]( , )/ 1 ( , ) ( , ) .NN F N NG dn eV T dV A Z B Z d ∞ −∞ = ε − + ε − ε ε∫ (3) Here GNN is the normal state conductance (or for the case of eV >> Δ); the electron energy ε is measured from the Fermi level of the superconductor; nF is the Fermi func- tion; the function AN(ε,Z) gives the probability of the AR and BN(ε,Z) corresponds to usual reflection. The conductance of the completely polarized channel GPS can be given in the form: ( )PSG V = [ ]( , )/ 1 ( , ) ( , )NF F P PG dn eV T dV A Z B Z d ∞ −∞ = ε − + ε − ε ε∫ (4) where GNF is the conductance of the contact for eV >> Δ, the parameter AP (ε,Z) describes the AR of polarized cur- rent, and BP (ε,Z) describes the usual reflection. If |ε| < Δ, the AR of completely polarized electrons is impossible and AP (ε,Z) = 0. The explicit expressions for the amplitudes AN (ε,Z), BN (ε,Z), AP (ε,Z), and BP (ε,Z), can be found in [14,27,56,57,61,63,64]. 3.2. Diffusive-type contacts For contacts with a normal state resistance RN < 100 Ω, the diffusive conductivity approximation is more appropri- ate. Accordingly, the electron momentum is no longer a good quantum number since l << d but with only elastic scattering in the contact region (i.e., d << lel, where lel is the inelastic-scattering length). This limit usually relates to N-SC contacts made by rubbing. Typically, it is assumed also that the superconductor remains in the clean limit (i.e., lS >> ξS) but, in addition to the N/SC interface, there is a wide layer of disordered ferromagnetic metal (adjacent to the contact area) where the electron mean free path is much shorter than the contact size. The total conductance of the contact in the diffusive approximation is given by Eqs. (2)–(4), where now all the amplitudes, AN (ε,Z), BN (ε,Z), AP (ε,Z), and BP (ε,Z), must be accordingly generalized. We will not write out here the extended expressions referring the reader to reports [13,18,24,27]. If not only the mean free path l is much less than the di- ameter, but the inelastic-scattering length, as well, d >> lel, we deal with the so-called thermal regime of conductance. There is both elastic and inelastic scattering in the contact region. In this case, Joule heating occurs in the contact region and causes a local increase in temperature. The in- terpretation of data obtained on such contacts is an ambi- guous task (see, e.g., discussion in [12,13]). 3.3. Contacts with spin-dependent transmission In spite of the conventional models success in fitting the experiments, several basic questions naturally arise for the SC-F contacts. For example, how to describe spin- dependent properties of the SC/F interface? Indeed, if the Fermi velocities for spin up and down electrons are differ- ent (vF↑ ≠ vF↑), then the effective parameter Z, and thus all other transmission coefficients, are also different for the majority- and minority – spin bands. The minimal model [65,66] which describes transport in the SC-F contacts and accounts for the spin-dependent transmission gives the following expressions for zero- temperature conductance: 1 2 2 , , (1 ) 4 ( / ) SF NN T T G G eV r r r r eV − ↑ ↓ ↑ ↓ ↑ ↓ = ≤ Δ + − Δ (5a) 1 2 ( ) ( / ) , , [(1 ) (1 ) ( / )] SF NN T T T T T T eV G G eV r r r r eV − ↑ ↓ ↑ ↓ ↑ ↓ ↑ ↓ ↑ ↓ + + − β Δ = ≥ Δ − + + β Δ (5b) where Tσ (σ = ↑ or↓ ) is the effective transmission coeffi- cient for electrons with spin up and down, respectively; rσ = (1 – Tσ) 1/2, β(eV/Δ) = [1 – (Δ/eV)2]1/2. The normal state conductance is read as GNN = (T↑ + T↓)/4, and the current polarization is defied as PC = (T↑ – T↓)/(T↑ + T↓). The con- tact conductivity at T ≠ 0 is given by Eq. (3) with GSF (5a), (5b) replacing the factor [ ]1 ( , ) ( , ) .N NA Z B Z+ ε − ε The main approximation of this model is the assumption that we can describe the point contact with a single pair of transmission coefficients, T↑ and T↓, which is finally justi- fied by the agreement with the experiment. Summarizing, the effects of the interface structure, spe- cific details of the contact preparation method, etc. should also be taken into account when finding the polarization of the ferromagnetic materials. In fact, one needs to test all models to obtain trustful value of spin polarization. For example, if the contact is in the diffusive limit, then fitting the data using the ballistic model may result in an underes- timated current polarization. For more detailed discussion of these questions we refer to Refs 12,13 and references therein. The examples of such fitting will be discussed in the next section. 4. Conventional half-metallic characteristics of point contacts We start with the results from samples that reveal con- ventional properties of sSC-HMF PCs. Conventional sce- nario predicts that in the extreme case of a completely spin polarized metal being in contact with a singlet s-wave SC (sSC) the AR is suppressed. Thus, a fingerprint of the conventional half-metallic characteristics of PC is a quite visible increase of the contact’s resistivity just after su- Point-contact Andreev-reflection spectroscopy of doped manganites: Charge carrier spin-polarization Low Temperature Physics/Fizika Nizkikh Temperatur, 2013, v. 39, No. 3 281 perconducting transition temperature of the electrode. In Fig. 1, representative low-temperature characteristics of Pb/La0.65Sr0.35MnO3 PC are shown. The I–V dependence is shown in the main panel, the top inset exhibits the tem- perature dependence of the contact’s resistance R(T), and the bottom inset illustrates the contact’s AR spectra, the dif- ferential conductance dI/dV. The quasiparticles energy gap of Pb is pointed in the bottom inset by arrows. (Further on, we simply denote the position of dI/dV minimum by Δ. For a PC with not too large Γ parameter, introduced by Dynes et al. [67], this value does not differ too much from the true energy gap [68].) Sharp increase of the contact’s resistivity just after TC(Pb) = 7.2 K is the main distinguishing feature of such type PCs. Below TC(Pb) we observe excess voltage and almost doubling of the contact’s resistivity. Reduction in the conductivity for the PC shown is about 50%. Similar results for Ca-doped manganite are shown in Fig. 2. Here temperature dependence of the Pb/La0.67Ca0.33MnO3 contact resistance is shown in the main panel. Inset in this figure illustrates the AR spectra for this PC. Figure 3 illustrates normalized conductivity (dI/dV)/(I/V) of some other contacts: MgB2/La0.65Ca0.35MnO3 (1) Pb/La0.7Sr0.3MnO3 (2), and Pb/La0.65Ca0.35MnO3 (3) (the graphs are shifted along the vertical axis for clarity); note that in case (3) the contact is in the thermal regime. Fitting the model predictions for different junction conduction to the experimental data (in Fig. 3, theory is solid line, expe- riment is symbols) one can restore the degree of charge carrier spin polarization for a given compound. The degree of spin polarization of charge carriers obtained by this Fig. 1. The current-voltage dependence of the Pb/La0.65Sr0.35MnO3 contact without visible superconducting proximity effect; T = = 4.2 K. Top inset: the temperature dependence of the contact’s resistance R(T); arrow indicates TC Pb = 7.2 K. Bottom inset: the contact’s Andreev-reflection spectra; T = 4.2 K; arrows indicate superconducting gap of Pb. Adopted from Ref. 29. 0.9 1.8 1.5 1.2 5 10 15 20 25 30 35 R , � T, K –5–10 6.0 0 5 10 5.5 5.0 4.5 4.0 3.5 3.0 d I/ d V , ar b . u n it s V, mV 2�Pb TC Pb –5–10 0 5 10 V, mV –5 –10 0 5 10 15 –15 I, m A Fig. 2. Temperature dependence of a resistance of Pb/La0.67Ca0.33MnO3 contact. Inset: Andreev-reflection spectra. Adopted from Ref. 28. 4 6 8 10 12 30 35 40 45 50 55 –4 –2 0 2 4 0.4 0.6 0.8 1.0 T = 4.2 K Voltage, mV T, K R , � d I/ d V , ar b . u n it s Fig. 3. Normalized conductivity (dI/dV)/(I/V) of MgB2/La0.65Ca0.35MnO3 (1) Pb/La0.7Sr0.3MnO3 (2), and Pb/La0.65Ca0.35MnO3 (3) PCs (the graphs are shifted along the vertical axis for clarity). In case (3), the contact is in the thermal regime (theory is solid line, experiment is symbols). The degree of the spin polarization of charge carriers obtained by fitting are PC ≈ 83, 78, and 65%, respectively. T = 4.2 K. Adopted from Ref. 27. –14 –7 0 7 14 0,3 0,6 0,9 1,2 V, mV 3 2 1 d I/ d V / I/ V ( ) V.N. Krivoruchko, A.I. D’yachenko, and V.Yu. Tarenkov 282 Low Temperature Physics/Fizika Nizkikh Temperatur, 2013, v. 39, No. 3 mode for the PCs shown in Fig. 3 is PC equals 83, 78, and 65%, respectively. The values of charge carrier spin polarization obtained for doped manganites are summarized in Table 1. In spite of some scattering in absolute magnitudes, in general, the data in the Table yields that the degree of spin polarization of the current can be unambiguously determined using the PCAR method. As already mentioned, the electron and the hole involved in the AR are coherently coupled. The phase-coherent elec- tron-hole conversion results in a nonzero pair amplitude in the F metal. For singlet pairing in the dirty limit, super- conducting coherence length is ξF ~ (DF/2πTCurie)1/2 and for manganites one obtains an extremely short distance ξF ~ 5–7 Ǻ. (Here DF is the diffusivity of the F metal; we choose = kB = 1.) Contribution of such a small region to the contact’s resistance is less than 1%. That is indeed observed for the most part of the samples [15,22–29,49– 55]. However, for some of the cases, the measured con- tact’s spectra revealed very distinct features which were interpreted as manifestation of an unconventional prox- imity effect. 5. Proximity affected contacts, “anomalous” Andreev reflection Typically, the normal state resistivity of proximity af- fected contacts is larger than that for the contacts without proximity effect (PE). Following the conventional physics [see discussion below Eq. (1)], the conductivity of these contacts should be close to the ballistic type and, accor- dingly, the spectroscopic characteristics would be more perfect than those shown in Figs. 1–3. Yet, the characteris- tics observed were fundamentally different. Figure 4 illustrates representative low-temperature cha- racteristics of Pb/La0.7Sr0.3MnO3 proximity affected PC. As in Fig. 1, the I–V dependence is shown in the main pan- el, the top inset exhibits the temperature dependence of the contact’s resistance R(T), and the bottom inset illustrates the contact’s AR spectra. In contrast to the data in Fig. 1, quite visible decrease of the contact’s resistivity is seen just after superconducting transition of a SC tip. Decrease of the PC resistivity is the main distinguishing feature of proximity affected contacts. Also, in contrast to the data in Fig. 1, below TC(Pb) = 7.2 K we now observe excess cur- rent and almost doubling of the contact’s conductivity. Figure 5 illustrates the properties of another Pb/La0.7Sr0.3MnO3 PC. In comparison with the PC in Fig. 4, the normal-state resistivity of this contact is about six times larger; however, it demonstrates all specific fea- tures of a proximity affected contact, namely: excess cur- rent, anomalous AR spectra, proximity induced gap, Δtr, Table 1. Degree of spin polarization deduced for doped manga- nites. Technique: SPT — spin-polarized tunneling; SPPh — spin- polarized photoemission; BC — band calculation; PCAR — point- contact Andreev reflection; STS — scanning tunneling spectrosco- py; ARPES — angle-resolved photoemission spectroscopy Compound Spin polarization Technique, Refs. La0.67Sr0.33MnO3 36% BC, [38] La0.67Sr0.33MnO3 72% SPT, [53] La0.67Sr0.33MnO3 54% SPT, [51] La0.67Sr0.33MnO3 81% SPT, [52] La0.7Sr0.3MnO3 100% SPPh, [5] La0.7Sr0.3MnO3 78% PCAR, [27] La0.7Sr0.3MnO3 58%–92% PCAR, [22] La0.7Sr0.3MnO3 78% PCAR, [23] La0.6Sr0.4MnO3 83% PCAR, [23] La0.7Sr0.3MnO3 78% PCAR, [15] La0.66Sr0.34MnO3 100% BC, [42] La0.6Sr0.4MnO3 100% ARPES, [48] La2/3Sr1/3MnO3 90% SPT, [55] La2/3Sr1/3MnO3 78%-82% PCAR, [24] La0.7Sr0.3MnO3 100% SPPh, [46] La0.7Sr0.3MnO3 83% PCAR, [23] La1–xCaxMnO3 100% BC, [39] La0.65Ca0.35MnO3 65%–83% PCAR, [27] La2/3Ca1/3MnO3 36% BC, [38] La0.7Ca0.3MnO3 100% STS, [44] La1–xCaxMnO3 80% BC, [43] La0.75Ca0.25MnO3 100% BC, [40] La0.7Ca0.3MnO3 86% SPT, [54] La0.7Ce0.3MnO3 100% SPPh, [47] La0.7Ce0.3MnO3 100% BC, [41] Fig. 4. The current-voltage dependence of the proximity affected Pb/La0.7Sr0.3MnO3 contact; T = 4.2 K. Top inset: the temperature dependence of the contact’s resistance R(T). Bottom inset: the contact’s Andreev-reflection spectra at T = 4.2 K. Reproduced form Ref. 29. –30 –20 –10 0 10 20 30 –6 –3 0 3 6 0.15 0.20 0.25 0.30 0.35 0.40 –30 –15 0 15 304 6 8 10 12 14 16 18 2.5 3.0 3.5 4.0 4.5 5.0 R , � d I/ d V , ar b . u n it s V, mV V, mV I, m A T, K Point-contact Andreev-reflection spectroscopy of doped manganites: Charge carrier spin-polarization Low Temperature Physics/Fizika Nizkikh Temperatur, 2013, v. 39, No. 3 283 which is much larger than that for Pb. In the bottom inset, the evolution of the AR spectra with temperature is shown. These data directly prove that the anomalous behavior of the junction is due to the superconducting state of Pb. In the top inset, the temperature dependence of proximity induced quasiparticle gap is shown. In Figs. 6 and 7, the normalized conductance spectra for some additional proximity affected PCs of Pb or MgB2 with films of Ca- and Sr-doped manganites are shown. For low voltage, the fine structure of the AR spectra is directly visible. This structure is the so-called subharmonic gap structure (SGS) and manifests itself in a set of downward peaks in the differential conductance that are pointed by labeled arrows in Figs. 6 and 7. According to contemporary models (see, for example, Refs. 69–71 and references therein), for SC–N–SC weak links or short constrictions SC–c–SC between two super- conductors, the differential conductance dI/dV drops fairly abruptly due to multiple ARs. These conductance drops appear at voltage that correlates with the energy of quasi- particle gaps divided by integers. The voltages at which the conductance SGS appears (roughly, because the boundary conditions at interface are also important [71]) are: eVn = = Δ1/n, eVm = Δ2/m, and eVl = (Δ1 + Δ2)/(2l + 1), where the integers (l, n, m) are restricted depending on the energy gap ratio (Δ1/Δ2). What is important for us here is that the resonances can be observed only if both electrodes are superconductors. So, the observation of the SGS (Figs. 6 and 7) is a strong argument in favor of the fact that the manganites are in a superconducting state with actual gap independent on SC partner. From the experimental results in Figs. 4–7, we extract that the proximity induced single-particle gap at the Pb/La0.7Sr0.3MnO3 and Pb/La0.65Ca0.35MnO3 interface is as large as Δtr ≈ 18–20 meV. Note that the detected gap Δtr is much larger than that of Pb or of MgB2; for these SCs we have at T = 4.2 K: Δ = 1.41 meV for Pb, and for MgB2 two superconducting energy gaps (the σ and π gaps) with Δπ = 2.3 meV and Δσ = 7.1 meV for MgB2 [72]. Knowing energy gap for the SC tip, the subharmonic gap structure can be classified as shown in Table 2. To conclude on this section, we summarize the main re- sults detected for proximity affected PCs (Figs. 4–7). Firstly, such principal fact as spectacular drop of the con- tact’s resistance with the onset of the SC tip superconduc- tivity has been observed. Secondly, the subharmonic gap resonances due to multiple AR are directly visible. Thirdly, in proximity affected PCs, the magnitude of a proximity induced gap is much larger than that of the SC tip and may be as large as Δtr ≈ 18–20 meV. These facts strongly sug- gest that both electrodes are in a superconducting state with independent gaps. All these anomalies are observed only in the superconducting state of the tip. Fig. 5. The current-voltage dependence of the proximity affected Pb/La0.7Sr0.3MnO3 contact; T = 4.2 K. Top inset: the evolution of the proximity induced quasiparticle gap Δtr with temperature. Bottom inset: the temperature dependence of the contact’s An- dreev-reflection spectra. The curves are shifted for clarity. Re- produced form Ref. 29. –48 –36 –24 –12 0 12 24 36 48 –1.5 –1.0 –0.5 0 0.5 1.0 1.5 4.0 4.5 5.0 5.5 6.0 6.5 0 2 4 6 8 10 12 0.98 1.00 1.02 1.04 1.06 –20 –10 0 10 20 T = 4.2 K 5.4 K 6.0 K 6.3 K 6.7 K d I/ d V , ar b . u n it s V, mV V, mV I, m A T, K 2 , m eV � tr Fig. 6. Normalized conductance spectra for proximity affected Pb/La0.65Ca0.35MnO3 point contacts (CP#2 and CP#4), and MgB2/La0.65Ca0.35MnO3 contact (CMg#7). The curves are shifted for clarity. The arrows indicate the energies of the subhar- monic gap resonances for CP#2 contact. See Table 3 for classifica- tion of the resonances; Δtr is the apparent proximity induced single- particle gap of the La0.65Ca0.35MnO3. Reproduced form Ref. 28. –60 –30 0 30 60 3 4 5 6 7 8 9 2�tr d d c b c b aCPb#2 CPb#4 CMg#7 V, mV d I/ d V , ar b . u n it s V.N. Krivoruchko, A.I. D’yachenko, and V.Yu. Tarenkov 284 Low Temperature Physics/Fizika Nizkikh Temperatur, 2013, v. 39, No. 3 To proceed further, let us briefly discuss models ex- tending the conventional concepts of interplay between superconductivity and ferromagnetism. 6. Triplet transport in proximity affected sSC–HMF point contacts Conventional model of PE predicts that proximity in- duced superconductivity decays rapidly (a few nanome- ters) inside the F layer owing to the incompatible nature of singlet superconductivity and ferromagnetic order, and thus proximity induced superconductivity of the F metal can be neglected [59]. This expectation was indeed con- firmed in various materials and geometries. On the other hand, an increasing number of experimental facts [73-82] present clear evidences that a simple physical interpreta- tion of the PE, reading that the Cooper pairs are broken by strong exchange field in the F layer, is in reality too sim- plistic, and an extension of the existing concepts of inter- play between superconductivity and ferromagnetism is needed. Long-ranged PE has been observed in a variety of ferromagnetic materials, including wires [74,74], bi- and multilayers [75,81], half metallic CrO2 [78] rare earth met- als with helical magnetic structure [77], etc. Concerning the manganites, note that much earlier Kasai et al., investigated current-voltage characteris- tics of YBCO/La1–xCaxMnOz/YBCO [83] and YBCO/ La1–xSrxMnOz/YBCO [84] layered junctions (here YBCO stands for YBa2Cu3Oy). Surprisingly, supercurrent was observed through magnetic barrier as thick as 300 nm for junctions with La1–xCaxMnOz and 200 nm for junctions with La1–xCaxMnOz. That is for barrier’s thicknesses much larger than a distance one may expect based on con- ventional proximity effect. Yet, this phenomenon occurred only for manganese oxides with x = 0.3–0.4. The authors suggested the results may be due to a novel proximity ef- fect between YBCO and doped manganites. Further neu- tron measurements on YPrBaCuO/La0.7Ca0.3MnO3 multi- layers [85] suggest a possibility of inducing spin-triplet superconducting phase in manganite layers, which could be the source of long-range proximity effect observed in La0.7Ca0.3MnO3/YBCO multilayers [76,86], trilayers [87], and bilayers [89]. From the theoretical viewpoint, a hybrid system of an F with a uniform exchange field in a metallic contact with a SC is well understood and the PE may be described by taking into account the splitting of electronic bands of op- posite spins [59]. The situation becomes more complicated if the magnetic structure is inhomogeneous. Theories [89– 93] predict the appearance of the long-range unconven- tional PE if there is spatial variation of magnetization (or exchange field) at the SC/F interface. Particularly, the triplet components of correlations need to be taken into consideration with a characteristic coherence length of ξF = (DF/2πT)1/2 that can be as large as ~ 100 nm at low temperatures. Before giving a qualitative explanation of the results detected on PE PCs (Figs. 4–7), we summarize the physics of proximity effect at spin-active sSC/F interface which looks as follows (for details see, e.g., [92,93] and refe- rences therein). 6.1. Triplet pairing in sSC/F heterostructures As it is well known, superconducting correlations are quantified by the anomalous Green’s function (see, e.g., [94]): Gαβ(r1,τ1;r2,τ2) = <TτΨα(r1,τ1)Ψβ(r2,τ2)>. Here all notations are conventional: the field operators Ψα(ri,τi) hold anticommutation relations, α and β are spin indexes, Tτ is an ordering operator for imaginary time τ, etc. The Fig. 7. Those as in Fig. 6 for proximity affected Pb/La0.7Sr0.3MnO3 point contact at 4.2 K. Reproduced form Ref. 29. –40 –20 0 20 40 0.008 0.012 0.016 0.020 0.024 0.028 0.032 f f d dc c bb a a 2 �tr V, mV d I/ d V , ar b . u n it s LSMO-Pb Table 2. The voltage corresponding to the SGS in Figs. 6 and 7, point contacts Pb–(La,Ca)MnO3 and Pb–(La,Sr)MnO3, respec- tively; here ΔPb ≈ 1.4 meV, Δtr ≈ ΔLCMO ≈ ΔLSMO ≈ 18–20 meV Label (Fig. 6, CPb#2) Voltage a ΔPb b ΔLCMO/2 c 2ΔLCMO/3 d ΔLCMO Label (Fig. 7) Voltage a ΔPb b (ΔPb + ΔLSMO)/5 c ΔLSMO/3 d ΔLSMO/2 f ΔLSMO Point-contact Andreev-reflection spectroscopy of doped manganites: Charge carrier spin-polarization Low Temperature Physics/Fizika Nizkikh Temperatur, 2013, v. 39, No. 3 285 Pauli principle requires that this function changes sign when the two particles are interchanged. For Fourier trans- formed Green’s function its reads: Gαβ(p,ωn) = – Gβα(-p,-ωn). (6) The symmetry restriction on Eq. (6) in spin, S, momen- tum, p, and Matsubara frequency, ωn = (2n + 1)πT, can be satisfied in four different ways (see, e.g., Table 1 in Ref. 92). It is convenient to separate the Green’s function Gαβ(p,ωn) into singlet and triplet components as Gαβ(p,ωn) = Gs(p,ωn) (iσy)αβ + Gtr(p,ωn) (σiσy)αβ. (7) Here σ is the vector of the three Pauli matrices. The singlet spin matrix (iσy)αβ is odd under the interchange α ↔ β, while the three triplet matrices (σ iσy)αβ are even. To illustrate the physics, we consider a case of a super- conductor-weak ferromagnet heterostructure in the ballistic transport regime and spin-active interface scattering. For simplicity, it was also suggested that the proximity induced pairing amplitudes in the F are small. In this case, within quasi-classical approximation, the anomalous Green’s function follows the (linearized) Eilenberger equations (we follow Ref. 92): ( ) tr exc2 2 sgn( ) 2s F n nf iH f ↑↓ ↑↓ ∇ + ω = πΔ ω −v (8) ( ) tr exc2 2 s F n f iH f ↑↓ ↑↓ ∇ + ω = −v (9) ( ) tr ,2 0.F n f ↑↑ ↓↓ ∇ + ω =v (10) tr ( )f ↑↓ = ↑↓> + ↓↑> (m = 0) and tr , ( , )f ↑↑ ↓↓ = ↑↑> ↓↓> (m = 1) are normalized triplet pairing amplitudes, and ( )sf ↑↓ = ↑↓> − ↓↑> stands for the singlet one. The super- conducting gap Δ is nonzero in the SC, while the exchange field Hexc exists in the F. Note, the exchange field Hexc is presented only in the eq- uations for the pairing correlations involving two spin bands, sf ↑↓ and tr ,f ↑↓ and is absent for the pairing correlations involving one spin band tr , .f ↑↑ ↓↓ The eigenvalues of the sf ↑↓ and trf ↑↓ amplitudes for a given vF are exc2( i )/ .n n Fk H v± = ω ± Thus, both the singlet sf ↑↓ and the triplet trf ↑↓ amplitudes oscillate on the clean-limit mag- netic length scale ξF = vF/2Hexc, and decay exponentially on the length scale ξn = vF/2|ωn|; the latter is dominated by the lowest Matsubara frequency, ω0 = πT, and occurs on the thermal length scale ξT = vF/2πT. The equations for the tr ,f ↑↑ ↓↓ triplet pairing amplitudes do not contain the ex- change field, and these components are monotonic decaying functions on the thermal length scale ξT = vF/2πT. Yet, the presence of the tr , ( , )f ↑↑ ↓↓ = ↑↑> ↓↓> compo- nents requires spin-active interface scattering, i.e., appro- priate boundary conditions. Indeed, the conversion process between the singlet and equal-spin triplet supercurrents is governed by two important phenomena taking place at the SC/F interface: (i) a spin mixing and (ii) a spin-flip scatter- ing. Spin mixing is the result of scattering phase difference θ that electrons with opposite spin acquire when scattered (reflected or transmitted) from an interface [95]: i i( ) ( e e )θ − θ↑↓> − ↓↑> => ↑↓> − ↓↑> = ( )cos( ) ( )sin( ).i= ↑↓> − ↓↑> θ + ↑↓> + ↓↑> θ It results from difference in orientations between magneti- zation of the F film and the spin of quasiparticle, or differ- ences in the wave-vector mismatches for spin up and spin down quasiparticles, etc. It is a robust and ubiquitous fea- ture for interfaces involving spin-polarized ferromagnets. However, triplet spin state ( )↑↓> + ↓↑> can transform into equal-spin pair states ( , )↑↑> ↓↓> only if there are spin-flip processes at the interfaces or if quantization direc- tion changes, or both. Its origin depends on microscopic magnetic state at the SC/F interface, character of local magnetic moments coupling with itinerant electrons, etc., and even varies from sample to sample. But, the exact mi- croscopic origin of spin-flip processes at the interface is important only for the effective interface scattering matrix [92,93,96,97] and not for superconducting phenomena, since Cooper pairs are of the size of the coherence length ξS which is much larger than the atomic scale. Thus, due to spin mixing at the interfaces, a spin triplet (S = 1, m = 0) amplitude, tr ,f ↑↓ is created and extends from the interface about the length ξT = vF/2πT into the F layer oscillating on very short length scale ξF = vF/2Hexc (in typical cases exchange field Hexc is much larger than TC). At the same time, triplet pairing correlations with equal spin pairs (S = 1, m = +1 or m = –1) are also induced (due to spin-flip processes) in the F layer smoothly decaying on the length scale ξT = vF/2πT. It is worthy to emphasize that it is only the m = 0 triplet component that is coupled via the spin-active boundary condition to the equal-spin m = 1 pairing amplitudes in the half-metal. The singlet compo- nent in the s-wave superconductor, ,sf ↑↓ being invariant under rotations around any quantization axis, is not directly involved in the creation of the triplet m = 1 pairing ampli- tudes in the half-metal. Considering the simplest spin-active model in which the scattering matrix is independent on the momentum parallel to the interface, for small constant transmission and small spin-mixing one can obtain analytical expressions for the amplitudes. Referring the reader for details to Ref. 92, for the singlet and triplet amplitudes with zero spin projection we have in the region ξF << x << ξT: ( )sf l ↑↓ = 0 sin ( / ) cos( / ) cos cos exp( / ), / / F F T F F x x f x x x ⎡ ⎤ξ ξ = − θ + α − ξ⎢ ⎥ξ ξ⎣ ⎦ V.N. Krivoruchko, A.I. D’yachenko, and V.Yu. Tarenkov 286 Low Temperature Physics/Fizika Nizkikh Temperatur, 2013, v. 39, No. 3 tr ( )f l ↑↓ = 0 cos( / ) sin( / ) cos cos sin exp( / ). / / F F T F F x x if x x x ⎡ ⎤ξ ξ = − θ − α θ − ξ⎢ ⎥ξ ξ⎣ ⎦ Here l is Legendre polynomials index; α is an angle be- tween a different spin-quantization axis and θ stands for the spin-mixing angle of spin-active interface; the ampli- tude f0 is determined by the related matrix elements [92]. The triplet amplitudes with nonzero spin projection in this region is approximately constant and the asymptotic beha- vior is reached if x >> ξT, and is of the form tr 0, ( ) ~ sin exp( / ).T Tf l if x x↑↑ ↓↓ ξ θ − ξ In Fig. 8 the first partial wave (l = 0) of the singlet ( )sf x ↑↓ and triplet, tr ( )f x ↑↓ and tr , ( ),f x ↑↑ ↓↓ pairing amplitudes [ ( 0),sf l = ( 0)tzf l = and ( 0),tf l⊥ = respectively] spatial behaviors are shown in clean sSC/F structures at tempera- ture near the superconducting critical temperature for the lowest Matsubara frequency, ω0 = πT. The higher or- der partial waves (l ≠ 0) look very similar and have simi- lar amplitudes. As can be seen in Fig. 8(a), in the region ξF << x << ξ0 = ξT the correlation functions ( )sf x ↑↓ and tr ( )f x ↑↓ decay like 1/x and are rapidly reduced by a factor ξF/ξT (= 0.01 for the data in Fig. 8). On the same scale the triplet correlation function with non-zero spin projection tr , ( )f x ↑↑ ↓↓ varies smoothly and very slowly. In a much larger distance, x >> ξT, Fig. 8(b), all components continue to decay according to (1/x) exp(-x/ξT). However, the mag- nitudes of ( )sf x ↑↓ and tr ( )f x ↑↓ are considerably reduced compared with tr , ( )f x ↑↑ ↓↓ before this region is reached. (For more details see Ref. 92.) The picture changes a little bit in the case of a fully spin polarized ferromagnet. Now the conversion of singlet pairs into triplet ones takes place entirely within the singlet SC. To be definite, we concentrate on an sSC/HMF structure suggesting the dirty limit for both metals. Due to spin mixing at the interfaces, a spin triplet (S = 1, m = 0) amplitude tr ( )f x ↑↓ is created. This results in a boun- dary layer with coexisting singlet ( )sf x ↑↓ and triplet tr ( )f x ↑↓ amplitudes near the interface which extend about a coherence length ξS = (DS/2πT)1/2 into the SC. Triplet com- ponents tr , ( )f x ↑↑ ↓↓ are generated if spin-flip centers are present in the interface region. Being created, the equal-spin proximity-induced amplitudes decay slowly in the half- metallic region on the thermal length scale ξT = (DF/2πT)1/2. The magnitude of the triplet correlations at the interface is proportional to that of the singlet amplitude at the inter- face, and both are insensitive to impurity scattering. (More discussion is given in Ref. 92.) Naturally, the physical mechanisms applicable for sSC/HMF heterostructures remain in force for the point- contact geometry, as well. The impact of spin-mixing and spin-flip AR processes at sSC/HMF interface on PC spectra was theoretically studied just recently in [93,96–99]. It was found that spin-active interface and spin-flip AR can be re- sponsible for the long-range triplet proximity effect. Refer- ring the reader to these reports for a detailed discussion of the problem, we only note here that the authors of Refs. [93,98,99] have proposed an alternative interpretation of the PCAR experimental results that goes beyond the de Jong and Beenakker theory [14]. This alternative interpretation is based on a realistic suggestion that, as already mentioned in Sec. 3.3, it is reasonable to expect that the scattering proper- ties of quasiparticles depend on their spin. In particular, the model utilizes the spin mixing angle θ. It was shown that scattering phase may play an important role in AR process at interfaces to strong ferromagnets. Indeed, the scattering phase difference θ, that quasiparticles incident from the sSC acquire upon being reflected at sSC/HMF interface, induces a triplet Cooper pairs, tr ,f ↑↓ which leads to enhanced subgap conductance. This allows for an alternative interpretation of Fig. 8. Pair correlation functions in the ferromagnet of a ballistic s-wave superconductor–weak spin-polarized ferromagnet junc- tion. Adopted from Ref. 92. 0 1 0 2 0 x/�F –0.2 –0.1 0 0.1 0.2 0.3 0.4 Re ( = 0)f ls Im ftz( = 0)l Im ft�( = 0)l (a) 0 50 100 –0.2 –0.1 0 0.1 0.2 0.3 0.4 x = �0 (b) x/�F Point-contact Andreev-reflection spectroscopy of doped manganites: Charge carrier spin-polarization Low Temperature Physics/Fizika Nizkikh Temperatur, 2013, v. 39, No. 3 287 the PCAR spectra, which in this picture can be turned by the fitting parameter θ [98,99]. It is worth to notice that, when no tunneling potential is present, an inverse magnetic proximity effect [100,101] may affect the sSC/F interface transport properties, too. 6.2. Triplet supercurrent in sSC-manganites PCs Let us now go back to the results in Figs. 4–7. If the con- tact’s size is not small, d > ξS, the PE can be important. For proximity affected contacts, it was suggested [25,26,28,29] that the conditions for an unconventional PE are fulfilled. Indeed, the main condition for spin triplet pairing to be induced at sSC/F interface is the “spin active” interface, i.e., the ability of the sSC/F interface to convert a singlet pair into a triplet one. For manganites, several theoretical models and numerous experimental data, including such local probing as nuclear magnetic resonance [102,103], point that nanoscale nonhomogeneity is an intrinsic feature of these compounds. Another characteristic important for our discussion is that, due to strong Hund’s interaction, spin disorder serves as strong spin scattering center for charge carriers. Accordingly, depending on the local magnetic non- homogeneity at the sSC/doped manganite boundary, the manganites surface causes coherent equal spin p-wave even frequency pairing correlations, which spread over large dis- tance into the manganite’s bulk [25,26,28,29]. However, the induction of pairing correlations in the normal region is not enough for the realization of multiple AR. As already mentioned, the observation of the sub- harmonic structure requires the existence of actual gaps in both superconducting electrodes. Thus, experimental find- ing of the SGS point that the proximity induced supercon- ducting state of manganites possesses an intrinsic super- conducting gap [28,29]. Following the physics described, in Fig. 9, the spatial structure of the current through the proximity affected con- tact is shown. The figure explains mutual conversion of the currents along the contact. In fact, due to a long-range PE, we deal here with a charge transport through an sSC–tSC– HMF heterostructure. Namely, there is a region at the sSC/HMF interface where a conversion from spin singlet pairs into spin triplet pairs takes place. The equal spin trip- let supercurrent flows through the HMF, while the singlet part is completely blocked. The sum of the singlet and trip- let currents is constant, obeying the continuity equation. At the boundary of superconducting and normal phases of the manganite a spin polarized supercurrent is continued as a quasiparticle current jq due to the usual AR mechanism. In.deed, at both sides of the tSC/HMF interface the charge current is spin polarized and there is no restriction (because of spin) on the AR. As a result, excess current and doubl- ing of the normal-state conductance have to be observed. The region where transformation of spin singlet pair in- to spin triplet pairs (and vise verse) takes place is shown in Fig. 10. Figure 10(a) illustrates the so-called “semiconduc- tor picture” of the proximity affected PC. Figure 10(b) explains the mechanism of conversion between spin singlet and spin triplet pairs due to multiple ARs. At the sSC/tSC interface we deal, in fact, with a weak link (or short constric- tion) between two different superconductors. The “weak link” here is a region where both singlet and triplet pairing amplitudes are suppressed. In the semiclassical picture, for a given voltage V < Δ/e across a weak link a quasiparticle accelerated from Fermi surface suffers n ~ Δ/eV ARs until it reaches the top of the pair potential well. In the particular case shown in Fig. 10(b), an incoming electron/hole of a given spin subband and under the energy gap ΔsSC cannot enter in the triplet superconducting electrode. It is spin flipped and then Andreev reflected (spin-flip AR process) as a hole/electron back to the sSC, simultaneously adding a triplet Cooper pair to the condensate in the tSC. This hole/electron is spin flipped and then is reflected by An- dreev mechanism as an electron/hole back to the tSC, si- multaneously adding a singlet Cooper pair to the conden- sate in the sSC. For a given voltage across the sSC/tSC interface a quasiparticle undergoes n ~ ΔsSC/eV, ΔtSC/eV, or m ~ (ΔsSC + ΔtSC)/eV reflections (depending on elec- trodes and energy it starts) until it reaches the top of the pair potential well. As was already indicated, within this physics the subharmonic peaks shown in Figs. 6 and 7 can be specified as it is given in Table 2. 6.3. Latent superconductivity of doped manganites Let us now make some suggestions concerning the ori- gin of the quasiparticle gap Δtr the magnitude of which cannot be explained in terms of conventional theory of proximity effect. In mean field BCS-Eliashberg theories with Δ(r) = = |ΔMF(r)|exp{φ(r)}, the characteristic energy scale re- sponsible for the global transition temperature TC is the superconducting energy gap |ΔMF(r)|. This silently implies that the spatial variation in |ΔMF(r)| is small, and that glob- al phase coherence temperature Tφ is larger than (or equal to) TC. However, for a system with small superfluid densi- Fig. 9. Spatial structure of the current through the proximity af- fected sSC/half-metallic manganite contact. The x axis is directed perpendicular to the sSC/HMF interface that is at x = 0; the HMF is placed in the region x > 0, while the sSC is located at x < 0. Proximity affected region LPE is much longer than the supercon- ducting coherence length ξT = (DF/2πT)1/2 and LPE >> ξT. Repro- duced form Ref. 28. LPE >> ξT j q spin-polarizedj S singlet j S triplet singlet SC normal phase x = 0 x j S triplet Half-metal V.N. Krivoruchko, A.I. D’yachenko, and V.Yu. Tarenkov 288 Low Temperature Physics/Fizika Nizkikh Temperatur, 2013, v. 39, No. 3 ty (bad metal) the spatial variations in the mean-field value |ΔMF(r)| could be large. As a consequence, due to large spa- tial variations, fluctuation effects become crucial in the re- gions where Δ(r) is small. Of these, the most important are the thermal fluctuations in the phase of the order parameter φ(r). In this case, the fluctuations in the phase of the order parameter in mesoscopic “islands” prevent the long-range superconductivity, i.e., the global critical temperature TC is determined by the global phase coherency, whereas the pair condensate could exist well above TC [104,105]. The important consequence of the presence of the Cooper pair fluctuation above TC is an appearance of the so-called pseudogap [104,106,107] i.e., decreasing of the one-electron density of state near the Fermi level. In par- ticular, according to one point of view [104], in the pseu- dogap state high-TC cuprates could be considered as an unconventional metal, i.e., as a SC that has lost its phase rigidity due to phase fluctuations. As already mentioned, doped manganites are bad metals. Also, a large pseudogap is indeed detected in numerous experiments on manganites [33–36] and it may be suggested that at least a portion of the observed pseudogap value is due to pairing without global phase coherency. Precursor diamagnetism above TC provides additional arguments for survival of the pair con- densate well above TC in cuprates [108,109]. However, for manganites, this kind of response may be strongly sup- pressed by ferromagnetic order of the localized moments and spin-triplet state of the condensate. As a likely hypothesis, the authors [28,29] suggested that the results obtained on proximity affected contacts are the observation of a new type of superconducting proximi- ty effect which follows the scenario of a proximity induced superconducting order parameter phase stiffness. Namely, the manganites are thermodynamically very close to a trip- let p-wave superconducting state and, at low temperature, the local triplet superconducting fluctuations with pairing energy Δtr intrinsically exist (“latent” high-TC supercon- ductivity of doped manganites). Being proximity coupled with singlet superconductor, the phase coherency of a su- perconducting state is restored. To verify this phase-disordering scenario for anomalous superconductivity of proximity affected PCs, the authors of a communication [110] prepared and studied normal and superconducting properties of the MgB2 — (nano) La0.67Sr0.33MnO3 (MgB:LSMO) composite. The key idea was to obtain such a composite, where proximity affected HMF/SC interfaces govern superconducting properties of the bulk sample. Fortunately, the idea was successful: the bulk samples of MgB:LSMO (nano)composite demonstrate direct evidences for unconventional superconductivity. Superconductivity of MgB:LSMO samples with 3:1 and 4:1 weight ratio has been observed with large, up to 20 K, critical temperature. A few features have been detected for bulk samples’ characteristics which, most probably, can hardly be explained within the framework of the conven- tional percolation model. Using the point-contact spectros- copy, three distinct quasiparticle energy gaps Δ1(π), Δ2(σ), and Δtr are clearly revealed. Two of these gaps were identi- fied as enhanced gaps in the quasiparticle spectrum of the MgB2 in the composite; the third gap Δtr was the same as those earlier detected in PCs of (La,Sr)MnO3 and (La,Ca)MnO3 with Pb or MgB2. A noteworthy argument was the temperature behavior of the Δtr gap which did not follow the BCS dependence. The Andreev-reflection spec- troscopy on PCs between the samples and half-metallic La0.65Ca0.35MnO3 electrode provides an additional evi- dence in favor of an unconventional superconducting state in the MgB:LSMO composite. The results obtained assert upon the new type of super- conducting proximity effect which provides for the phase- coherency stiffness. At low temperature in a half-metallic ferromagnetic state of (La,Sr)MnO3, a phase incoherent Fig. 10. Semiconductor picture of proximity affected sSC/half- metallic manganite contact. Weak link here is a region where both singlet and triplet pairing interactions are suppressed (a). Trajectory of a quasiparticle that is accelerated out of the conden- sate by the electric field suffering multiple Andreev reflections. In the case of singlet and triplet superconducting electrodes every Andreev reflection is foregone with a spin-flip scattering. Spin- flip Andreev-reflection processes are illustrated by lines with stars (a). Reproduced form Ref. 29. HMF x = 0 x Weak link Triplet SC Normal phase Weak link ΔtS ΔsS HMF’s surface region e e e e h h tSCsSC sSC LPE T>> ξ ΔtS ΔsS Point-contact Andreev-reflection spectroscopy of doped manganites: Charge carrier spin-polarization Low Temperature Physics/Fizika Nizkikh Temperatur, 2013, v. 39, No. 3 289 superconductivity (local triplet pairing condensate) exists. However, though the local gap amplitude is large, there is no phase stiffness and the system is incapable to display long-range superconducting response. Being proximity coupled to MgB2, the long-range coherency is restored. Inversely, the manganite in superconducting state with large energy pairing, due to proximity effect, enhances the MgB2 superconducting state. That is, here we deal with some kind of “mutual” proximity effect. Naturally, at this stage of investigation, other possible explanations and/or mechanisms, which could be related to the physics of proximity affected PCs of manganites should also be discussed. It was predicted that in sSC/F structures, the so-called odd frequency pairing could take place [89,90,111]. In this case, the Cooper pair wave func- tion is symmetric under exchange of spatial- and spin- coordinates but antisymmetric under exchange of time- coordinates. The study of such pairing in sSC/F junction was addressed by a number of authors over the last years [111]. However, if the superconducting correlations are odd in frequency, a pairing interaction has to be additional- ly frequency dependent (due to strong retardation effect) in order to have a nonzero intrinsic gap in the ferromagnetic metal (see details in Ref. 92). Nonlocal or crossed AR in which an electron from one magnetic domain is Andreev reflected as a hole into oppo- sitely polarized domain while a pair is transmitted into a superconductor [112–114] is, in principle, possible. How- ever, in order for the data on proximity affected junctions to be in agreement with the crossed AR mechanism, the total current through the contact has to be unpolarized. In may be if the portion of domains with opposite magnetiza- tion is exactly equal. It seems improbable that in all the proximity affected PCs the portion of domains with oppo- site magnetization is exactly equal. A conversion of spin-less Cooper pair into spin pola- rized Cooper pair and vice versa is also possible due to absorption (respectively, emission) of a magnon [115]. If this mechanism is governing, the junction’s current-voltage characteristic has to be, at low temperatures, asymmetric with respect to the base voltage. The I(V) characteristics of all PCs explored are symmetric, and thus it is hard to sug- gest that the magnon assisted mechanism controls the charge transport in PCs. A giant proximity effect (a logarithmic dependence of the junction critical temperature on the junction width) was predicted for a tunnel junction of two SCs with the barrier formed by a SC that has lost its phase rigidity due to phase fluctuations [116]. We think that to a certain extent this scenario of proximity effect could be relevant to our case. Extended discussion, including more exotic explana- tions and/or mechanisms, which could be related to the physics of proximity affected PCs of manganites the reader can find in Ref. [28,29]. 7. Conclusions The interplay between superconductivity and spin- polarized materials has potential applications in the emerg- ing field of spin-electronics. Specifically, the so-called superconducting spintronics [3,117,118] is among the most attractive subjects of spintronics, and requires a class of superconducting materials with spin-polarized transport, which would necessarily have to be triplet. The controlled production of triplet supercurrents will open several direc- tions for possible applications. With the availability of ful- ly polarized triplet supercurrents, spin-dependent quantum- coherence phenomena will make implementation of differ- ent spintronics devices a reality. Superconducting spintron- ics devices are appealing since they introduce in a natural way the elements of nonlocality, entanglement, and quan- tum coherence, all of which are crucial, e.g., for quantum computing. Already existing data [73–82] demonstrate that an effective source of spin-polarized supercurrent can be designed using nanohybrids of metallic ferromagnets and a conventional s-wave SCs. Out-of-the-mainstream is the idea that, at low temperature, incoherent superconducting fluctuations are essentially sustained in half-metallic man- ganites and, in proximity affected region, the singlet super- conductor establishes phase coherence of the spin- polarized superconducting state of the manganites [28,29]. If it is a success, a promising source of spin-polarized su- percurrent can be designed using nano-junctions of man- ganites and a conventional s-wave SC. Systematic character and repeatability of a number of principal experimental facts suggest that some general physical phenomena have been observed in transport prop- erties of proximity affected singlet superconductor-half- metallic manganite contacts. Whether the origin of these features can be traced to a thermodynamic state of manga- nites with local triplet superconducting fluctuations, is a matter of further investigations. Further experiments are definitely needed in order to prove (or disprove) this scena- rio and understand the mechanism causing the local triplet pairing in doped manganites. We hope that our short re- view stimulates further experimental and theoretical works on this subject. 1. J.F. Gregg, I. Petej, and C. Dennis, J. Phys. D: Appl. Phys. 35, R121 (2002). 2. A.M. Bratkovsky, Rep. Prog. Phys. 71, 026502 (2008). 3. Y.C. Tao and J.G. Hu, J. Appl. Phys. 107, 041101 (2010). 4. G.-X. Miao, M. Münzenberg, and J.S. Moodera, Rep. Prog. Phys. 74, 036501 (2011). 5. J.-H. Park, E. Vescovo, H.-J. Kim, C. Kwon, R. Ramesh, and T. Venkatesan, Nature (London) 392, 794 (1998). 6. J.M.D. Coey, J.J. Versluijs, and M. Venkatesan, J. Phys. D: Appl. Phys. 35, 2457 (2002). 7. R. Meservey and P.M. Tedrow, Phys. Rep. 238, 173 (1994). 8. A.F. Andreev, Sov. Phys. JETP 19, 1228 (1964). 9. I.K. Yanson, Sov. Phys. JETP 39, 506 (1974). V.N. Krivoruchko, A.I. D’yachenko, and V.Yu. Tarenkov 290 Low Temperature Physics/Fizika Nizkikh Temperatur, 2013, v. 39, No. 3 10. I.K. Yanson, Fiz. Nizk. Temp. 9, 676 (1983) [Sov. J. Low Temp. Phys. 9, 343 (1983)]. 11. I.K. Yanson, and O.I. Shklyarevskii, Fiz. Nizk. Temp. 12, 885 (1986) [Sov. J. Low Temp. Phys. 12, 509 (1986)]. 12. Y.G. Naidyuk and I.K. Yanson, Point-Contact Spectroscopy, Springer (2005). 13. D. Daghero and R.S. Gonelli, Supercond. Sci. Technol. 23, 043001 (2010). 14. M.J.M. de Jong and C.W. Beenakker, Phys. Rev. Lett. 74, 1657 (1995). 15. R.J. Soulen, Jr., J.M. Byers, M.S. Osofsky, B. Nadgorny, T. Ambrose, S.F. Cheng, P.R. Broussard, C.T. Tanaka, J. Nowak, J.S. Moodera, A. Barry, and J.M.D. Coey, Science 282, 85 (1998). 16. S.K. Upadhyay, A. Palanisami, R.N. Louie, and R.A. Buhrman, Phys. Rev. Lett. 81, 3247 (1998). 17. G.J. Strijkers, Y. Ji, F.Y. Yang, C.L. Chien, and J.M. Byers, Phys. Rev. B 63, 104510 (2001). 18. B. Nadgorny, R.J. Soulen, Jr., M.S. Osofsky, I.I. Mazin, G. Laprade, R.J.M. van de Veerdonk, A.A. Smits, S.F. Cheng, E.F. Skelton, and S.B. Qadri, Phys. Rev. B 61, R3788 (2000). 19. R.J. Soulen, Jr., M.S. Osofsky, B. Nadgorny, T. Ambrose, P. Broussard, S.F. Cheng, J. Byers, C.T. Tanaka, J. Nowack, J.S. Moodera, G. Laprade, A. Barry, and M.D. Coey, J. Appl. Phys. 85, 4589 (1999). 20. Y. Ji, G.J. Strijkers, F.Y. Yang, and C.L. Chien, Phys. Rev. B 64, 224425 (2001). 21. J.S. Parker, S.M. Watts, P.G. Ivanov, and P. Xiong, Phys. Rev. Lett. 88, 196601 (2002). 22. B. Nadgorny, I.I. Mazin, M. Osofsky, R.J. Soulen, Jr., P. Broussard, R.M. Stroud, D.J. Singh, V.G. Harris, A. Arsenov, and Y. Mukovskii, Phys. Rev. B 63, 184433 (2001). 23. Y. Ji, C.L. Chien, Y. Tomioka, and Y. Tokura, Phys. Rev. B 66, 012410 (2002). 24. G.T. Woods, R.J. Soulen, Jr., I. Mazin, B. Nadgorny, M.S. Osofsky, J. Sanders, H. Srikanth, W.E. Egelhoff, and R. Datla, Phys. Rev. B 70, 054416 (2004). 25. V.N. Krivoruchko, V.Yu. Tarenkov, A.I. D’yachenko, and V.N. Varyukhin, Europhys. Lett. 75, 294 (2006). 26. A.I. D’yachenko, V.N. Krivoruchko, and V.Yu. Tarenkov, Fiz. Nizk. Temp. 32, 1085 (2006) [Low Temp. Phys. 32, 824 (2006)]. 27. A.I. D’yachenko, V.A. D’yachenko, V.Yu. Tarenkov, and V.N. Krivoruchko, Phys. Solid State 48, 432 (2006). 28. V.N. Krivoruchko and V.Yu. Tarenkov, Phys. Rev. B 75, 214508 (2007). 29. V.N. Krivoruchko and V.Yu. Tarenkov, Phys. Rev. B 78, 054522 (2008). 30. I. Leonov, A.N. Yaresko, V.N. Antonov, and V.I. Anisimov, Phys. Rev. B 74, 165117 (2006). 31. Yu.S. Dedkov, U. Rüdiger, and G. Güntherodt, Phys. Rev. B 65, 064417 (2002). 32. P.Y. Song, J.F. Wang, C.P. Chen, H. Deng and Y.D. Li, J. Appl. Phys. 100, 044314 (2006). 33. E. Dagotto, T. Hotta, and A. Moreo, Phys. Rep. 344, 1 (2001). 34. M.B. Salamon and M. Jaime, Rev. Mod. Phys. 73, 583 (2001). 35. V.M. Loktev and Yu.G. Pogorelov, Fiz. Nizk. Temp. 26, 231 (2000) [Low Temp. Phys. 26, 171 (2000)]. 36. J.M.D. Coey, M. Viret, and S. von Molnar, Adv. Phys. 48, 167 (1999). 37. V.N. Krivoruchko and S.I. Khartsev, Fiz. Nizk. Temp. 24, 1070 (1998) [Low Temp. Phys. 24, 803 (1998)]. 38. W.E. Pickett and D.J. Singh, Phys. Rev. B 53, 1146 (1996). 39. S. Satpathy, Z.S. Popovic, and F.R. Vukajlovic, Phys. Rev. Lett. 76, 960 (1996). 40. W.J. Chang, J.Y. Tsai, H.-T. Jeng, J.-Y. Lin, Kenneth Y.-J. Zhang, H.L. Liu, J.M. Lee, J.M. Chen, K.H. Wu, T.M. Uen, Y.S. Gou, and J.Y. Juang, Phys. Rev. B 72, 132410 (2005). 41. Q. Zhang and W. Zhang, Phys. Rev. B 68, 134449 (2000). 42. Ch. Ma, Zh. Yang, and S. Picozzi, J. Phys. Condens. Matter. 8, 7717 (2006). 43. I. Golosov, Phys. Rev. Lett. 104, 207207 (2010). 44. J.Y.T. Wei, N.-C. Yeh, and R.P. Vasquez, Phys. Rev. Lett. 79, 5150 (1997). 45. Y. Okimoto, T. Katsufuji, T. Ishikawa, A. Urushibara, T. Arima, and Y. Tokura, Phys. Rev. Lett. 75, 109 (1995). 46. R. Bertacco, M. Portalupi, M. Marcon, L. Duó, F. Ciccacci, M. Bowen, J.P. Contour, and A. Barthélémy, J. Magn. Magn. Mater. 242–245, 710 (2002). 47. S.W. Han, J.-S. Kang, K.H. Kim, J.D. Lee, J.H. Kim, S.C. Wi, C. Mitra, P. Raychaudhuri, S. Wirth, K.J. Kim, B.S. Kim, J.I. Jeong, S.K. Kwon, and B.I. Min, Phys. Rev. B 69, 104406 (2004). 48. A. Chikamatsu, H. Wadati, H. Kumigashira, M. Oshima, A. Fujimori, N. Hamada, T. Ohnishi, M. Lippmaa, K. Ono, M. Kawasaki, and H. Koinuma, Phys. Rev. B 73, 195105 (2006). 49. H.Y. Hwang, S-W. Cheong, N.P. Ong, and B. Batlogg, Phys. Rev. Lett. 77, 2041 (1996). 50. J.Z. Sun, W.J. Gallagher, P.R. Duncombe, L. Krusin- Elbaum, R.A. Altman, A. Gupta, Yu Lu, G.Q. Gong, and Gang Xiao, Appl. Rev. Lett. 69, 3266 (1996). 51. Y. Lu, X.W. Li, G.Q. Gong, G. Xiao, A. Gupta, P. Lecoeur, J.Z. Sun, Y.Y. Wang and V.P. Dravid, Phys. Rev. B 54, R8357 (1996). 52. J. Z. Sun, L. Krusin-Elbaum, P.R. Duncombe, A. Gupta, and R.B. Laibowitz, Appl. Phys. Lett. 70, 1769 (1997). 53. D.C. Worledge and T.H. Geballe, Appl. Phys. Lett. 76, 900 (2000). 54. Moon-Ho Jo, N.D. Mathur, N.K. Todd, and M.G. Blamire, Phys. Rev. B 61, 14905 (2000). 55. M. Bowen, M. Bibes, A. Barthélémy, J.-P. Contour, A. Anane, Y. Lemaître, and A. Fert, Appl. Phys. Lett. 82, 233 (2003). 56. I.I. Mazin, Phys. Rev. Lett. 83, 1427 (1999). 57. I.I. Mazin, A.A. Golubov, and B. Nadgorny, J. Appl. Phys. 89, 7576 (2001). 58. G. Wexler, Proc. Phys. Soc. 89, 927 (1966). 59. A.I. Buzdin, Rev. Mod. Phys. 77, 935 (2005). Point-contact Andreev-reflection spectroscopy of doped manganites: Charge carrier spin-polarization Low Temperature Physics/Fizika Nizkikh Temperatur, 2013, v. 39, No. 3 291 60. A.V. Zaitsev, Zh. Eksp. Teor. Fiz. 78, 221 (1980). 61. G.E. Blonder, M. Tinkham, and T.M. Klapwijk, Phys. Rev. B 25, 4515 (1982). 62. Yu.V. Sharvin, Sov. Phys. JETP 21, 655 (1965). 63. C.H. Kant, O. Kurnosikov, A.T. Filip, P. LeClair, H.J.M. Swagten, and W.J.M. de Jonge, Phys. Rev. B 66, 212403 (2002). 64. P. Chalsani, S.K. Upadhyay, O. Ozatay, and R.A. Buhrman, Phys. Rev. B 75, 094417 (2007). 65. A. Martín-Rodero, A. Levy Yeyati, and J.C. Cuevas, Physica C 353, 67 (2001). 66. F. Pérez-Willard, J.C. Cuevas, C. Surgers, P. Pfundstein, J. Kopu, M. Eschrig, and H.V. Lohneysen, Phys. Rev. B 69, 140502 (2004). 67. R.C. Dynes, V. Narayanamurti, and J.P. Garno, Phys. Rev. Lett. 41, 1509 (1978). 68. E.L. Wolf, Principles of Electron Tunneling Spectroscopy, Oxford University Press, New York (1985). 69. U. Gunsenheimer and A.D. Zaikin, Phys. Rev. B 50, 6317 (1994). 70. M. Hurd, S. Datta, and P.F. Bagwell, Phys. Rev. B 54, 6557 (1996). 71. B.A. Aminov, A.A. Golubov, and M.Yu. Kupriyanov, Phys. Rev. B 53, 365 (1996). 72. X.X. Xi, Rep. Prog. Phys. 71, 116501 (2008). 73. M. Giroud, H. Courtois, K. Hasselbach, D. Mailly, and B. Pannetier, Phys. Rev. B 58, R11872 (1998). 74. V.T. Petrashov, I.A. Sosnin, I. Cox, A. Parsons, and C. Troadec, Phys. Rev. Lett. 83, 3281 (1999). 75. M.D. Lawrence and N. Giordano, J. Phys. Condens. Matter 11, 1089 (1999). 76. V. Peňa, Z. Sefrioui, D. Arias, C. Leon, J. Santamaria, M. Varela, S.J. Pennycook, and J.L. Martinez, Phys. Rev. B 69, 224502 (2004). 77. I. Sosnin, H. Cho, V.T. Petrashov, and A.F. Volkov, Phys. Rev. Lett. 96, 157002 (2006). 78. R.S. Keizer, S.T.B. Goennenwein, T.M. Klapwijk, G. Maio, G. Xiao, and A. Gupta, Nature (London) 439, 825 (2006). 79. T.S. Khaire, M.A. Khasawneh, W.P. Pratt, Jr., and N.O. Birge, Phys. Rev. Lett. 104, 137002 (2010). 80. J. Wang, M. Singh, M. Tian, N. Kumar, B. Liu, C. Shi, J.K. Jain, N. Samarth, T.E. Mallouk, and M.H.W. Chan, Nature Phys. 6, 389 (2010). 81. Y. Kalcheim, O. Millo, M. Egilmez, J.W.A. Robinson, and M.G. Blamire, Phys. Rev. B 85, 104504 (2012). 82. C. Klose, T.S. Khaire, Yixing Wang, W.P. Pratt, Jr., N.O. Birge, B.J. McMorran, T.P. Ginley, J.A. Borchers, B.J. Kirby, B.B. Maranville, and J. Unguris, Phys. Rev. Lett. 108, 127002 (2012). 83. M. Kasai, T. Ohno, Y. Kanke, Y. Kozono, M. Hanazono, and Y. Sugita, Jpn. J. Appl. Phys. 29, L2219 (1990). 84. M. Kasai, Y. Kanke, T. Ohno, and Y. Kozono, J. Appl. Phys. 72, 5344 (1992). 85. J. Hoppler, J. Sthan, Ch. Nidermayer, V.K. Malik, H. Boyanfif, A.J. Drew, M. Rossle, A. Buzdin, G. Cristiani, H.U. Habermeier, B. Keimer, and C. Bernhard, Nature Mater. 8, 315 (2009). 86. Z. Sefrioui, D. Arias, V. Pena, J.E. Villegas, M. Varela, P. Prieto, C. Leon, J.L. Martinez, and J. Santamaria, Phys. Rev. B 67, 214511 (2003). 87. K. Dybko, K. Werner-Malento, P. Aleshkevych, M. Wojcik, M. Sawicki, and P. Przyslupski, Phys. Rev. B 80, 144504 (2009). 88. Y. Kalcheim, T. Kirzhner, G. Koren, and O. Millo, Phys. Rev. B 83, 064510 (2011). 89. A. Kadigrobov, R.I. Skekhter, and M. Jonson, Europhys. Lett. 54, 394 (2001). 90. F.S. Bergeret, A.F. Volkov, and K.B. Efetov, Phys. Rev. Lett. 86, 4096 (2001). 91. M. Eschrig, J. Kopu, J.C. Cuevas, and G. Schön, Phys. Rev. Lett. 90, 137003 (2003). 92. M. Eschrig and T. Löfwander, T. Champel, J. C. Cuevas, J. Kopu, and G. Schön, J. Low Temp. Phys. 147, 457 (2007). 93. R. Grein, T. Löfwander, G. Metalidis, and M. Eschrig, Phys. Rev. B 81, 094508 (2010). 94. A.A. Abrikosov, L.P. Gorkov, and I.E. Dzyaloshinski, Methods of Quantum Field Theory in Statistical Physics, Prentice-Hall, Englewood Cliffs NJ (1963). 95. T. Tukuyasu, J.A. Sauls, D. Rainer, Phys. Rev. B 38, 8823 (1988). 96. J.N. Kupferschmidt and P.W. Brouwer, Phys. Rev. B 83, 014512 (2011). 97. F.B. Wilken and P.W. Brouwer, Phys. Rev. B 85, 134531 (2012). 98. T. Löfwander, R. Grein, and M. Eschrig, Phys. Rev. Lett. 105, 207001 (2010). 99. S. Piano, R. Grein, C.J. Mellor, K. Výborný, R. Campion, M. Wang, M. Eschrig, and B.L. Gallagher, Phys. Rev. B 83, 081305(R) (2011). 100. V.N. Krivoruchko and E.A. Koshina, Phys. Rev.B 66, 014521 (2002). 101. F.S. Bergeret, A.F. Volkov, and K.B. Efetov, Phys. Rev. B 69, 174504 (2004). 102. M.M. Savosta, V.N. Krivoruchko, I.A. Danielenko, V.Y. Tarenkov, T.E. Konstantinova, A.V. Borodin, and V.N. Varyukhin, Phys. Rev. B 69, 024413 (2004). 103. A.S. Mazur, V.N. Krivoruchko, and I.A. Danilenko, Fiz. Nizk. Temp. 33, 1227 (2007) [Low Temp. Phys. 33, 931 (2007)]. 104. V.J. Emery and S.A. Kivelson, Phys. Rev. Lett. 74, 3253 (1995). 105. E. Berg, D. Orgad, and S.A. Kivelson, Phys. Rev. B 78, 094509 (2008). 106. V.M. Loktev, R.M. Quick, and S.G. Sharapov, Phys. Rep. 349, 1 (2001). 107. M. Norman, D. Pines, and C. Kallin, Adv. Phys. 54, 715 (2005). 108. Lu Li, Y. Wang, S. Omiya, S. Ono, Y. Ando, G.D. Gu, and N.P. Ong, Phys. Rev. B 81, 054510 (2010). 109. J. Mosqueira, J.D. Dancausa, and F. Vidal, Phys. Rev. B 84, 174518 (2011). V.N. Krivoruchko, A.I. D’yachenko, and V.Yu. Tarenkov 292 Low Temperature Physics/Fizika Nizkikh Temperatur, 2013, v. 39, No. 3 110. V.N. Krivoruchko and V.Y. Tarenkov, Phys. Rev. B 86, 104502 (2012). 111. F.S. Bergeret, A.F. Volkov, and K.B. Efetov, Rev. Mod. Phys. 77, 1321 (2005). 112. G. Deutscher and D. Feinberg, Appl. Phys. Lett. 76, 487 (2000). 113. D. Beckmann, H.B. Weber, and H. v. Lőhneysen, Phys. Rev. Lett. 93, 197003 (2004). 114. P. Aronov and G. Koren, Phys. Rev. B 72, 184515 (2005). 115. G. Tkachov, E. McCann, and V. Fal’ko, Phys. Rev. B 65, 024519 (2001). 116. D. Marchand, L. Covaci, M. Berciu, and M. Franz, Phys. Rev. Lett. 101, 097004 (2008). 117. S. Anders, M.G. Blamire, F.-Im. Buchholz, D.-G. Crété, R. Cristiano, P. Febvre, L. Fritzsch, A. Herr, E. Il’ichev, J. Kohlmann, J. Kunert, H.-G. Meyer, J. Niemeyer, T. Ortlepp, H. Rogalla, T. Schurig, M. Siegel, R. Stolz, E. Tarte, H.J.M. ter Brake, H. Toepfer, J.-C. Villegier, A.M. Zagoskin, and A.B. Zorin, Physica C 470, 2079 (2010). 118. M. Eschrig, Phys. Today 64, 43 (2011).