16.01.2015 Views

An infrared study of the Josephson vortex state in high-Tc cuprates

An infrared study of the Josephson vortex state in high-Tc cuprates

An infrared study of the Josephson vortex state in high-Tc cuprates

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

EUROPHYSICS LETTERS 1 January 2003<br />

Europhys. Lett., 61 (1), pp. 122–128 (2003)<br />

<strong>An</strong> <strong><strong>in</strong>frared</strong> <strong>study</strong> <strong>of</strong> <strong>the</strong> <strong>Josephson</strong> <strong>vortex</strong> <strong>state</strong><br />

<strong>in</strong> <strong>high</strong>-T c <strong>cuprates</strong><br />

S. V. Dordevic 1 , S. Komiya 2 , Y. <strong>An</strong>do 2 ,Y.J.Wang 3 and D. N. Basov 1<br />

1 Department <strong>of</strong> Physics, University <strong>of</strong> California at San Diego<br />

La Jolla, CA 92093, USA<br />

2 Central Research Institute <strong>of</strong> Electric Power Industry - Tokyo, Japan<br />

3 National High Magnetic Field Laboratory - Tallahassee, FL 3231, USA<br />

(received 19 June 2002; accepted <strong>in</strong> f<strong>in</strong>al form 4 October 2002)<br />

PACS. 74.72.Dn – La-based <strong>cuprates</strong>.<br />

PACS. 74.25.Gz – Optical properties.<br />

PACS. 74.50.+r – Proximity effects, weak l<strong>in</strong>ks, tunnel<strong>in</strong>g phenomena, and <strong>Josephson</strong> effects.<br />

Abstract. – Wereport<strong>the</strong>results<strong>of</strong><strong>the</strong>c-axis <strong><strong>in</strong>frared</strong> spectroscopy <strong>of</strong> La 2−xSr xCuO 4 <strong>in</strong><br />

<strong>high</strong> magnetic field oriented parallel to <strong>the</strong> CuO 2 planes. A significant suppression <strong>of</strong> <strong>the</strong><br />

superfluid density withmagnetic field ρ s(H) is observed for bothunderdoped (x =0.125) and<br />

overdoped (x =0.17) samples. We show that <strong>the</strong> exist<strong>in</strong>g <strong>the</strong>oretical models <strong>of</strong> <strong>the</strong> <strong>Josephson</strong><br />

<strong>vortex</strong> <strong>state</strong> fail to consistently describe <strong>the</strong> observed effects and discuss possible reasons for<br />

<strong>the</strong> discrepancies.<br />

Although <strong>the</strong> mechanism <strong>of</strong> <strong>high</strong>-T c superconductivity <strong>in</strong> <strong>cuprates</strong> still rema<strong>in</strong>s unresolved,<br />

significant progress has been made <strong>in</strong> <strong>the</strong> understand<strong>in</strong>g <strong>of</strong> <strong>the</strong> mixed <strong>state</strong> <strong>of</strong> <strong>the</strong>se systems<br />

[1]. Many subtle predictions <strong>of</strong> <strong>the</strong> <strong>the</strong>ories <strong>of</strong> <strong>the</strong> <strong>vortex</strong> <strong>state</strong> have been experimentally<br />

verified [1–18]. In particular, electrodynamics <strong>in</strong> relatively small magnetic fields (typically,<br />

H


S. V. Dordevic et al.: <strong>An</strong> <strong><strong>in</strong>frared</strong> <strong>study</strong> <strong>of</strong> <strong>the</strong> <strong>Josephson</strong> <strong>vortex</strong> <strong>state</strong> etc. 123<br />

Reflectance<br />

Reflectance<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

6 K<br />

17 K<br />

20 K<br />

23 K<br />

32 K<br />

0 T<br />

5 T<br />

10 T<br />

13 T<br />

15 T<br />

17 T<br />

H=0 T<br />

La 1.875<br />

Sr 0.125<br />

CuO 4<br />

T=6 K<br />

6 K<br />

21 K<br />

27 K<br />

30 K<br />

33 K<br />

36 K<br />

0 T<br />

3 T<br />

7 T<br />

10 T<br />

13 T<br />

17 T<br />

H=0 T<br />

La 1.83<br />

Sr 0.17<br />

CuO 4<br />

T=6 K<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

Reflectance Reflectance<br />

0.0<br />

0 20 40 60 80 100<br />

Frequency [cm -1 ]<br />

0 50 100 150 200 0.0<br />

Frequency [cm -1 ]<br />

Fig. 1 – Infrared data for LSCO crystals with x =0.125(leftpanels)andx =0.17 (right panels).<br />

The top panels show c-axis reflectance <strong>in</strong> zero field; <strong>the</strong> bottom panels show: R(ω) <strong>in</strong> <strong>high</strong> magnetic<br />

field.<br />

<strong>the</strong> relative errors <strong>of</strong> <strong>the</strong> field-<strong>in</strong>duced changes are less than 1%. The optical conductivity<br />

σ 1 (ω) +iσ 2 (ω) and <strong>the</strong> dielectric function ɛ 1 (ω) +iɛ 2 (ω) were calculated from R(ω) us<strong>in</strong>g<br />

Kramers-Kronig analysis.<br />

In <strong>the</strong> superconduct<strong>in</strong>g <strong>state</strong> <strong>the</strong> real part <strong>of</strong> <strong>the</strong> optical conductivity has two components:<br />

σ1 SC (ω) = ρs<br />

8<br />

δ(0) + σreg 1 (ω), <strong>the</strong> first one due to superconduct<strong>in</strong>g condensate and <strong>the</strong> second<br />

due to unpaired carriers below T c (a so-called regular contribution). The superfluid density<br />

ρ s is quantified with <strong>the</strong> plasma frequency ω s : ρ s = ωs 2 = c 2 /λ 2 c =4πe 2 n s /m ∗ related to<br />

<strong>the</strong> density <strong>of</strong> superconduct<strong>in</strong>g carriers n s and <strong>the</strong>ir mass m ∗ ; λ c is <strong>the</strong> penetration depth.<br />

A delta-function <strong>in</strong> σ1 SC (ω) gives rise to a term −(ω s /ω) 2 <strong>in</strong> ɛ 1 (ω). A common procedure <strong>of</strong><br />

extract<strong>in</strong>g ω s from ɛ 1 (ω) <strong>in</strong>volves fitt<strong>in</strong>g <strong>of</strong> <strong>the</strong> low-energy part <strong>of</strong> <strong>the</strong> spectrum with 1/ω 2 .The<br />

problem with this procedure is that it does not discrim<strong>in</strong>ate screen<strong>in</strong>g due to superconduct<strong>in</strong>g<br />

condensate from regular contribution. In order to fix this problem, we use <strong>the</strong> follow<strong>in</strong>g<br />

correction procedure:<br />

ɛ 1 (ω) − ɛ reg<br />

1 (ω) =−ω2 s<br />

ω 2 , (1)<br />

where ɛ reg<br />

1<br />

ɛ reg<br />

1<br />

(ω) is <strong>the</strong> regular contribution to <strong>the</strong> dielectric function. In order to calculate<br />

(ω), a KK-like transformation is employed:<br />

∫ ∞<br />

ɛ reg<br />

1 (ω) =1+2 ω ′ ɛ reg<br />

2 (ω′ )<br />

π 0 ω ′ 2<br />

− ω 2 dω′ , (2)<br />

+<br />

where ɛ reg<br />

2 (ω) is <strong>the</strong> regular contribution to <strong>the</strong> imag<strong>in</strong>ary part <strong>of</strong> <strong>the</strong> dielectric constant,<br />

i.e. after a δ(0)-function has been subtracted. The procedure described by eqs. (1) and (2)<br />

accounts for <strong>the</strong> contribution <strong>of</strong> unpaired carriers at T < T c , but also phonons, <strong>in</strong>terband<br />

transitions, magnons, and all o<strong>the</strong>r f<strong>in</strong>ite-energy excitations.<br />

Figure 1 displays <strong>the</strong> raw reflectance data. In zero field <strong>the</strong> 6 K reflectance <strong>of</strong> both LSCO<br />

crystals is characterized by a sharp plasma edge at <strong>the</strong> frequency ω s / √ ɛ ∞ (ɛ ∞ is <strong>the</strong> <strong>high</strong>-


124 EUROPHYSICS LETTERS<br />

Loss Function<br />

0.3<br />

0.2<br />

0.1<br />

0.0<br />

0.3<br />

0.2<br />

0.1<br />

6 K<br />

17 K<br />

20 K<br />

23 K<br />

26 K<br />

28 K<br />

32 K<br />

15 T<br />

17 T<br />

ω s [cm -1 ] 150<br />

0 T<br />

7 T<br />

13 T<br />

100<br />

50<br />

La 1.875<br />

Sr 0.125<br />

CuO 4<br />

0.0<br />

0 20 40 60<br />

0<br />

→<br />

H<br />

10 20 30<br />

T [K]<br />

Frequency [cm -1 ]<br />

.<br />

u<br />

H=0 T<br />

T=6 K<br />

La 1.83<br />

Sr 0.17<br />

CuO 4<br />

CuO 2<br />

10 20 30<br />

T [K]<br />

CuO 2<br />

→<br />

J<br />

300 ω s [cm-1 ]<br />

200<br />

100<br />

0<br />

6 K<br />

18 K<br />

21 K<br />

24 K<br />

27 K<br />

30 K<br />

33 K<br />

36 K<br />

0 T<br />

5 T<br />

10 T<br />

13 T<br />

17 T<br />

0.3<br />

0.2<br />

0.1<br />

0.0<br />

0.3<br />

0.2<br />

0.1<br />

0 30 60 90 120 0.0<br />

Frequency [cm -1 ]<br />

Loss Function<br />

Fig. 2 – The loss function Im(1/ɛ(ω)) reveals coupl<strong>in</strong>g to <strong>the</strong> longitud<strong>in</strong>al JPR mode. The two top<br />

panels show <strong>the</strong> temperature dependence <strong>of</strong> <strong>the</strong> loss function and <strong>the</strong> bottom ones <strong>the</strong> loss function<br />

<strong>in</strong> <strong>high</strong> magnetic field. The top <strong>in</strong>sets show <strong>the</strong> temperature dependence <strong>of</strong> <strong>the</strong> superfluid density<br />

and <strong>the</strong> bottom one sketches <strong>the</strong> geometry <strong>of</strong> a <strong>Josephson</strong> <strong>vortex</strong> experiment.<br />

frequency dielectric constant). This form <strong>of</strong> reflectance, which resembles a plasma edge <strong>in</strong><br />

normal metals, is due to <strong>the</strong> zero cross<strong>in</strong>g <strong>in</strong> <strong>the</strong> real part <strong>of</strong> <strong>the</strong> dielectric function, ɛ 1 (ω) =0,<br />

and is commonly referred to as <strong>the</strong> <strong>Josephson</strong> Plasma Resonance (JPR). As temperature <strong>in</strong>creases,<br />

<strong>the</strong> edge <strong>in</strong> R(ω) is smeared out and <strong>the</strong> m<strong>in</strong>imum shifts to lower frequency, <strong>in</strong>dicat<strong>in</strong>g<br />

suppression <strong>of</strong> <strong>the</strong> superfluid density. In <strong>the</strong> normal <strong>state</strong> <strong>the</strong> underdoped sample shows a very<br />

weak upturn <strong>of</strong> R(ω) asω → 0, <strong>in</strong>dicat<strong>in</strong>g a small metallic contribution at T>T c . The upturn<br />

is stronger <strong>in</strong> <strong>the</strong> x =0.17 crystal suggest<strong>in</strong>g that <strong>the</strong> far IR conductivity <strong>in</strong>creases with<br />

dop<strong>in</strong>g. Similar enhancement <strong>of</strong> <strong>the</strong> “metallic” trends <strong>in</strong> <strong>the</strong> <strong>in</strong>terlayer response is commonly<br />

found <strong>in</strong> o<strong>the</strong>r cuprate families. At <strong>the</strong> lowest measured temperature we found, us<strong>in</strong>g our<br />

new method described by eqs. (1) and (2), <strong>the</strong> follow<strong>in</strong>g values <strong>of</strong> <strong>the</strong> superfluid density at<br />

10 K: ω s = 160 cm −1 (λ c =9.7 µm) for 12.5% sample and ω s = 360 cm −1 (λ c =4.3 µm) for<br />

17% material. The <strong>in</strong>sets <strong>in</strong> <strong>the</strong> top panels <strong>of</strong> fig. 2 show <strong>the</strong> temperature dependence <strong>of</strong> <strong>the</strong><br />

superfluid density closely resembl<strong>in</strong>g <strong>the</strong> form expected for a d-wave superconductor.<br />

Application <strong>of</strong> a magnetic field has a strong impact on <strong>the</strong> JPR feature <strong>in</strong> both crystals<br />

(bottom panels <strong>of</strong> fig. 1): <strong>the</strong> plasma edges are smeared and <strong>the</strong> m<strong>in</strong>ima <strong>in</strong> R(ω, H) also shift<br />

to lower ω. The plasma edge shift <strong>in</strong> 17 T field is as strong as 35% <strong>in</strong> <strong>the</strong> x =0.125 sample<br />

and 10% <strong>in</strong> <strong>the</strong> x =0.17 crystal. This result is surpris<strong>in</strong>g, given <strong>the</strong> fact that <strong>the</strong> strongest<br />

field used <strong>in</strong> our experiments is still orders <strong>of</strong> magnitude smaller than <strong>the</strong> upper critical field<br />

H c2 for <strong>the</strong> H ‖ CuO 2 orientation.<br />

It is <strong>in</strong>structive to present <strong>the</strong> evolution <strong>of</strong> <strong>the</strong> superconduct<strong>in</strong>g condensate with H and T<br />

with <strong>the</strong> spectra <strong>of</strong> <strong>the</strong> loss function Im(1/ɛ(ω)) = ɛ 2 (ω)/(ɛ 2 1(ω)+ɛ 2 2(ω)) (fig. 2). At T ≪ T c ,<br />

<strong>the</strong> loss function is peaked at a frequency close, but not exactly equal to, <strong>the</strong> JPR, whereas <strong>the</strong><br />

width <strong>of</strong> <strong>the</strong> Im(1/ɛ(ω)) mode is proportional to <strong>the</strong> magnitude <strong>of</strong> ɛ 2 (ω, T < T c ). As temperature<br />

<strong>in</strong>creases <strong>the</strong> peak shifts to lower energies, <strong>in</strong>dicat<strong>in</strong>g suppression <strong>of</strong> <strong>the</strong> superfluid density.


S. V. Dordevic et al.: <strong>An</strong> <strong><strong>in</strong>frared</strong> <strong>study</strong> <strong>of</strong> <strong>the</strong> <strong>Josephson</strong> <strong>vortex</strong> <strong>state</strong> etc. 125<br />

Magnetic Field [T]<br />

H/H 0<br />

0 3 6 9 12 15 18 0.0 0.1 0.2 0.3 0.4 0.5<br />

100<br />

A<br />

B<br />

100<br />

ω s (H)/ω s (0 T) [%]<br />

90<br />

80<br />

70<br />

90<br />

80<br />

70<br />

ω s (H)/ω s (0 T) [%]<br />

60<br />

60<br />

100 C<br />

ω s (H)/ω s (0 T) [%]<br />

90<br />

80<br />

70<br />

60<br />

2∆ =70 cm -1<br />

2∆ =45 cm -1<br />

2∆ =35 cm -1<br />

La 1.83 Sr 0.17 CuO 4<br />

La 1.875 Sr 0.125 CuO 4<br />

0 2 4 6 8 10 12 14 16 18<br />

Magnetic Field [T]<br />

Fig. 3 – Change <strong>of</strong> <strong>the</strong> superfluid density with magnetic field ω s(H)/ω s(0) alone with <strong>the</strong> <strong>the</strong>oretical<br />

results obta<strong>in</strong>ed for <strong>the</strong> TKT model (panel A), Bulaevskii et al. model (panel B) and <strong>the</strong> scenario<br />

tak<strong>in</strong>g <strong>in</strong>to account <strong>the</strong> nodal Zeeman effect (panel C). While conventional models <strong>of</strong> <strong>the</strong> <strong>Josephson</strong><br />

<strong>vortex</strong> <strong>state</strong> are <strong>in</strong>consistent with <strong>the</strong> <strong>high</strong>-field data for <strong>the</strong> underdoped x =0.125 crystal (A and<br />

B), a plausible description can be achieved with<strong>in</strong> <strong>the</strong> picture discussed by Won et al. [20] us<strong>in</strong>g <strong>the</strong><br />

magnitude <strong>of</strong> <strong>the</strong> gap from <strong>the</strong> <strong>in</strong>-plane measurements for <strong>the</strong> same crystal [21].<br />

The behavior <strong>of</strong> <strong>the</strong> loss function <strong>in</strong> <strong>high</strong> magnetic field (fig. 2, bottom panels) is generally<br />

similar to <strong>the</strong> H = 0 data taken at f<strong>in</strong>ite temperature: <strong>the</strong> peak s<strong>of</strong>tens and its width is<br />

enhanced as <strong>the</strong> field <strong>in</strong>creases. Figure 3A quantifies <strong>the</strong> demise <strong>of</strong> ω s (extracted from ɛ(ω)<br />

spectra us<strong>in</strong>g eqs. (1) and (2)) <strong>in</strong> magnetic field. At 17 T <strong>the</strong> superfluid density is reduced by<br />

38% <strong>in</strong> <strong>the</strong> underdoped and 12% <strong>in</strong> <strong>the</strong> overdoped sample; <strong>the</strong>se values are <strong>in</strong> full agreement<br />

with <strong>the</strong> strength <strong>of</strong> <strong>the</strong> effect <strong>in</strong>ferred directly from raw data <strong>in</strong> fig. 1. We also note a<br />

qualitatively different form <strong>of</strong> <strong>the</strong> ω s (H) dependencies <strong>in</strong> <strong>the</strong> two samples. Two pr<strong>in</strong>cipal<br />

mechanisms <strong>of</strong> <strong>the</strong> superfluid density suppression <strong>in</strong> <strong>high</strong> magnetic field are: 1) direct pairbrak<strong>in</strong>g<br />

<strong>of</strong> <strong>the</strong> condensate and 2) dissipation associated with <strong>vortex</strong> dynamics. The former<br />

process is usually discarded <strong>in</strong> view <strong>of</strong> giant H c2 values <strong>in</strong> <strong>the</strong> H ‖ CuO 2 configuration.<br />

With<strong>in</strong> <strong>the</strong> latter picture <strong>the</strong> oscillat<strong>in</strong>g electric field with <strong>the</strong> E-vector along <strong>the</strong> c-axis leads<br />

to a transverse motion u <strong>of</strong> <strong>Josephson</strong> vortices [1] located between <strong>the</strong> CuO 2 planes <strong>in</strong> <strong>the</strong><br />

direction perpendicular to <strong>the</strong> field (see <strong>the</strong> bottom <strong>in</strong>set <strong>of</strong> fig. 2). The electrodynamics <strong>of</strong><br />

<strong>Josephson</strong> vortices has been thoroughly discussed <strong>in</strong> several papers [14–16]. Below we show<br />

that <strong>the</strong> behavior <strong>of</strong> JPR <strong>in</strong> LSCOcrystals cannot be fully understood with<strong>in</strong> <strong>the</strong> proposed<br />

models, especially <strong>in</strong> <strong>the</strong> underdoped sample.


126 EUROPHYSICS LETTERS<br />

Bulaevskii et al. have worked out <strong>the</strong> follow<strong>in</strong>g prediction for <strong>the</strong> field dependence <strong>of</strong> ω s [16]:<br />

[<br />

ω s (H) =ω 0 1 − π H<br />

ln H ]<br />

0<br />

. (3)<br />

8 H 0 H<br />

The effect <strong>of</strong> magnetic field is directly related to <strong>the</strong> strength <strong>of</strong> <strong>the</strong> coupl<strong>in</strong>g between <strong>the</strong><br />

CuO 2 layers that is parameterized through <strong>the</strong> characteristic field H 0 =Φ 0 /γs 2 . In this expression<br />

Φ 0 is <strong>the</strong> flux quantum, γ = λ c /λ ab (λ ab is <strong>the</strong> <strong>in</strong>-plane penetration depth) is <strong>the</strong><br />

anisotropy factor and s ≃ 1.32 nm is <strong>the</strong> <strong>in</strong>terlayer distance. We estimated H 0 =36.7T for<br />

<strong>the</strong> x =0.125 sample with λ ab =0.2 µm (ref. [22]) and H 0 =55Tfor<strong>the</strong>x =0.17 sample<br />

with λ ab =0.3 µm (ref. [22]). In fig. 3B we plot ω s (H)/ω s (0) as a function <strong>of</strong> H/H 0 along<br />

with <strong>the</strong> <strong>the</strong>oretical dependence (full l<strong>in</strong>e). The overall trends <strong>of</strong> ω s (H) are clearly different<br />

for <strong>the</strong> two crystals. In particular, <strong>the</strong> x =0.125 material reveals a dramatic reduction <strong>of</strong><br />

<strong>the</strong> condensate strength compared to <strong>the</strong> model prediction for H/H 0 > 0.3. For <strong>the</strong> x =0.17<br />

sample such <strong>high</strong> magnetic fields could not be achieved with <strong>the</strong> present experimental setup.<br />

We also attempted to describe ω s (H) with<strong>in</strong> a phenomenological scenario <strong>of</strong> <strong>vortex</strong> dynamics<br />

[14, 15]. Tachiki, Koyama and Takahashi (TKT) [15] obta<strong>in</strong>ed an explicit result for<br />

<strong>the</strong> complex dielectric function (at frequencies below <strong>the</strong> JPR) <strong>of</strong> a layered superconductor <strong>in</strong><br />

<strong>the</strong> mixed <strong>state</strong>:<br />

ωn<br />

2<br />

ω 2 + iγ sr ω + ω2 s<br />

ɛ(ω) =ɛ ∞ −<br />

ω 2<br />

1+ φ , (4)<br />

0 H<br />

4πλ 2 c κ p − iηω − Mω 2<br />

where M is <strong>the</strong> <strong>vortex</strong> <strong>in</strong>ertial mass, η is <strong>the</strong> viscous force coefficient and κ p is <strong>the</strong> <strong>vortex</strong><br />

p<strong>in</strong>n<strong>in</strong>g constant. In eq. (4), ɛ ∞ is <strong>the</strong> real part <strong>of</strong> <strong>the</strong> dielectric function above <strong>the</strong> plasmon<br />

and ω n and γ sr are <strong>the</strong> regular component (due to unpaired carriers below T c ) plasma frequency<br />

and scatter<strong>in</strong>g rate. Equation (4) can be regarded as a H ≠ 0 generalization <strong>of</strong> <strong>the</strong> well-known<br />

“two-fluid” model <strong>of</strong> superconductivity, commonly used <strong>in</strong> <strong>the</strong> microwave and IR frequency<br />

ranges [23]. In order to model <strong>the</strong> data <strong>in</strong> fig. 3 we extracted <strong>the</strong> ω s (H) behavior from eq. (4)<br />

by explor<strong>in</strong>g Re[ɛ(ω)] <strong>in</strong> <strong>the</strong> limit <strong>of</strong> ω → 0: ω 2 s = ω 2 × ɛ 1 (ω). Some <strong>of</strong> <strong>the</strong> parameters<br />

<strong>of</strong> <strong>the</strong> TKT equation needed to obta<strong>in</strong> this result were readily available from <strong>the</strong> fits <strong>of</strong><br />

R(ω, 6K,H = 0): ɛ ∞ = 27, ω n = 200 cm −1 (for <strong>the</strong> 12.5% sample) and 1100 cm −1 (for <strong>the</strong><br />

17% sample), γ sr = 5000 cm −1 . Similar to all previous spectroscopic works [5,6,24], we set <strong>the</strong><br />

<strong>vortex</strong> mass to zero <strong>in</strong> eq. (4). The viscous drag constant η can be calculated with<strong>in</strong> C<strong>of</strong>fey-<br />

Clem approach [14] yield<strong>in</strong>g η = 7 Pa cm and 28 Pa cm, for <strong>the</strong> underdoped and overdoped<br />

samples, respectively [25]. Therefore, we are left with <strong>the</strong> p<strong>in</strong>n<strong>in</strong>g force constant κ p as <strong>the</strong> only<br />

fitt<strong>in</strong>g parameter <strong>in</strong> eq. (4). As fig. 3A shows, <strong>the</strong> <strong>the</strong>ory gives a very good fit for <strong>the</strong> overdoped<br />

sample, with κ p = 6000–11000 Pa , a value comparable with that <strong>of</strong> nearly optimally doped<br />

YBa 2 Cu 3 O y [24] (<strong>in</strong> <strong>the</strong> same field configuration). For <strong>the</strong> underdoped sample on <strong>the</strong> o<strong>the</strong>r<br />

hand, we could not obta<strong>in</strong> a good fit for any value <strong>of</strong> κ p . In order to reproduce <strong>the</strong> overall<br />

depression <strong>of</strong> <strong>the</strong> superfluid density <strong>in</strong> <strong>the</strong> 17 T field, we have to adopt an unphysically small<br />

κ p = 150–200 Pa. We believe that such a vast difference <strong>in</strong> <strong>the</strong> magnitude <strong>of</strong> κ p between <strong>the</strong><br />

two samples is ano<strong>the</strong>r signal <strong>of</strong> <strong>in</strong>ability <strong>of</strong> <strong>the</strong> TKT scenario to account for <strong>the</strong> experimental<br />

situation at least <strong>in</strong> underdoped LSCOcrystals.<br />

One obvious dist<strong>in</strong>ction between <strong>the</strong> underdoped and overdoped samples is <strong>the</strong> different<br />

nature <strong>of</strong> <strong>the</strong> <strong>in</strong>terlayer transport. Earlier experiments [28] established that <strong>the</strong> x =0.16<br />

dop<strong>in</strong>g <strong>in</strong> <strong>the</strong> LSCOsystem separates <strong>the</strong> region <strong>of</strong> “<strong>in</strong>sulat<strong>in</strong>g” (for x0.16) ground <strong>state</strong>s. This result may have important implications for <strong>the</strong> nature <strong>of</strong> <strong>the</strong><br />

H ‖ CuO 2 -vortices. The x =0.125 crystal is likely to fall <strong>in</strong> <strong>the</strong> regime where a description


S. V. Dordevic et al.: <strong>An</strong> <strong><strong>in</strong>frared</strong> <strong>study</strong> <strong>of</strong> <strong>the</strong> <strong>Josephson</strong> <strong>vortex</strong> <strong>state</strong> etc. 127<br />

with<strong>in</strong> <strong>the</strong> formalism <strong>of</strong> <strong>Josephson</strong> vortices is applicable ow<strong>in</strong>g to <strong>the</strong> <strong>in</strong>sulat<strong>in</strong>g character <strong>of</strong> <strong>the</strong><br />

“medium” separat<strong>in</strong>g <strong>the</strong> CuO 2 planes. However, as dop<strong>in</strong>g progresses to <strong>the</strong> overdoped side<br />

eventually <strong>the</strong> H ‖ CuO 2 -vortices will evolve <strong>in</strong>to Abrikosov-type vortices hav<strong>in</strong>g a normal<br />

core. Therefore, our x =0.17 crystal may be closer to <strong>the</strong> regime where additional factors<br />

<strong>in</strong>volv<strong>in</strong>g complex character <strong>of</strong> <strong>the</strong> <strong>vortex</strong> cores <strong>in</strong> d-wave superconductors have to be taken<br />

<strong>in</strong>to consideration [13, 29, 30]. Surpris<strong>in</strong>gly, <strong>the</strong> conventional analysis (eqs. (3), (4)) obviously<br />

ignor<strong>in</strong>g <strong>the</strong> latter issues, is less problematic for <strong>the</strong> x =0.17 crystal but fails for <strong>the</strong> x =0.125<br />

sample which, based on its “<strong>in</strong>sulat<strong>in</strong>g” resistivity at T → 0, is more likely to comply with<br />

eqs. (3), (4). In particular, <strong>the</strong> TKT analysis yields an unexpectedly small p<strong>in</strong>n<strong>in</strong>g constant<br />

which is hard to reconcile with <strong>the</strong> prom<strong>in</strong>ence <strong>of</strong> <strong>the</strong> <strong>in</strong>tr<strong>in</strong>sic <strong>in</strong>homogeneities <strong>in</strong> underdoped<br />

<strong>cuprates</strong>, s<strong>in</strong>ce <strong>in</strong>homogeneities would normally promote <strong>vortex</strong> p<strong>in</strong><strong>in</strong>nig.<br />

When search<strong>in</strong>g for reasons for <strong>the</strong> <strong>in</strong>ability <strong>of</strong> eqs. (3), (4) to consistently describe <strong>the</strong><br />

anomalous sensitivity <strong>of</strong> <strong>the</strong> superfluid response <strong>in</strong> underdoped LSCOto magnetic field, it<br />

is prudent to revisit <strong>the</strong> assumptions <strong>of</strong> <strong>the</strong>se models. Indeed, eqs. (3), (4) are valid for<br />

<strong>Josephson</strong> coupl<strong>in</strong>g between uniform s-wave superconductors; validity <strong>of</strong> both assumptions<br />

for <strong>high</strong>-T c materials is questionable. One implication <strong>of</strong> <strong>the</strong> d-wave order parameter (firmly<br />

established for <strong>cuprates</strong>) is that <strong>the</strong> Zeeman energy associated with <strong>the</strong> H ‖ CuO 2 field<br />

can no longer be disregarded for <strong>state</strong>s close to <strong>the</strong> node [31]. The impact <strong>of</strong> <strong>the</strong> Zeeman<br />

field is especially important for <strong>the</strong> x =0.125 phase, s<strong>in</strong>ce at this particular composition <strong>the</strong><br />

magnitude <strong>of</strong> <strong>the</strong> <strong>in</strong>-plane superconduct<strong>in</strong>g gap is anomalously low [21]. The field-dependence<br />

√<strong>of</strong> <strong>the</strong> c-axis superfluid density with<strong>in</strong> this scenario has <strong>the</strong> follow<strong>in</strong>g form: ωs 2 (H)/ωs 2 (0) =<br />

1 − (H/∆)2 [20] and is displayed <strong>in</strong> fig. 3C for different magnitudes <strong>of</strong> <strong>the</strong> energy gap<br />

2∆. Interest<strong>in</strong>gly, <strong>the</strong> functional form <strong>of</strong> ω s (H)/ω s (H = 0) is reproduced by this calculation.<br />

The Zeeman reduction <strong>of</strong> <strong>the</strong> condensate strength occurs <strong>in</strong> concert with o<strong>the</strong>r dissipation<br />

mechanisms and accounts for at least 10–15% depression <strong>of</strong> <strong>the</strong> ω s (17 T) value. <strong>An</strong> additional<br />

factor pert<strong>in</strong>ent to <strong>the</strong> anomalous field response <strong>of</strong> <strong>the</strong> underdoped crystals may be connected<br />

to spatial non-uniformities <strong>of</strong> superconductivity with<strong>in</strong> <strong>the</strong> CuO 2 planes revealed by a variety<br />

<strong>of</strong> experimental methods [32]. It seems plausible that <strong>the</strong> magnetic field will <strong>in</strong>fluence coupl<strong>in</strong>g<br />

between <strong>the</strong>se dissimilar regions <strong>in</strong> <strong>the</strong> CuO 2 plane thus lead<strong>in</strong>g to <strong>the</strong> field dependence <strong>of</strong><br />

<strong>the</strong> <strong>in</strong>-plane superconduct<strong>in</strong>g parameters, such as λ ab <strong>in</strong> eq. (3).<br />

Models <strong>of</strong> <strong>the</strong> superfluid density suppression discussed above <strong>in</strong>volve only a reduction <strong>of</strong><br />

<strong>the</strong> carrier density n s . However, <strong>the</strong> c-axis response <strong>of</strong> <strong>cuprates</strong> also reveals changes <strong>of</strong> <strong>the</strong><br />

effective mass m ∗ or <strong>of</strong> <strong>the</strong> k<strong>in</strong>etic energy at T


128 EUROPHYSICS LETTERS<br />

∗∗∗<br />

We thank D. P. Arovas, L. Bulaevskii, Ch. Helm, J. R. Clem, M. C<strong>of</strong>fey, A. E.<br />

Koshelev, T. Koyama and A. J. Millis for useful discussions. The research is supported<br />

by NSF, DoE, <strong>the</strong> Research Corporation and <strong>the</strong> State <strong>of</strong> Florida.<br />

REFERENCES<br />

[1] Blatter G. et al., Rev. Mod. Phys., 66 (1994) 1125.<br />

[2] Yeh N.-C., Phys. Rev. B, 43 (1991) 523.<br />

[3] Fe<strong>in</strong>berg D. et al., Phys. Rev. Lett., 65 (1990) 919.<br />

[4] Doyle R. A. et al., Phys. Rev. Lett., 71 (1993) 4241.<br />

[5] Parks B. et al., Phys. Rev. Lett., 74 (1995) 3265.<br />

[6] Gerrits A. M. et al., Phys. Rev. B, 51 (1995) 12049.<br />

[7] Dulic D. et al., Phys. Rev. Lett., 86 (2001) 4660.<br />

[8] Pimenov A. et al., Phys. Rev. Lett., 87 (2001) 177003.<br />

[9] Pesetski A. A. et al., Phys. Rev. B, 62 (2000) 11826.<br />

[10] Choi E.-J. et al., Phys. Rev. B, 49 (1994) 13271.<br />

[11] Koshelev A. E. et al., Phys. Rev. Lett., 81 (1998) 902.<br />

[12] Matsuda Y. et al., Phys. Rev. Lett., 75 (1995) 4512.<br />

[13] Lake B. et al., Science, 291 (2001) 1759.<br />

[14] C<strong>of</strong>fey M. W. et al., Phys. Rev. Lett., 67 (1991) 386.<br />

[15] Tachiki M. et al., Phys. Rev. B, 50 (1994) 7065.<br />

[16] Bulaevskii L. N. et al., Phys. Rev. Lett., 74 (1995) 801.<br />

[17] Lihn H.-T. S. et al., Phys. Rev. B, 56 (1997) 5559.<br />

[18] Kakeya I. et al., cond-mat-0111094.<br />

[19] <strong>An</strong>do Y. et al., Phys. Rev. Lett., 87 (2001) 017001.<br />

[20] Won H. et al., Physica B, 281-282 (2000) 944.<br />

[21] Dumm M. et al., Phys. Rev. Lett., 88 (2002) 147003. IR measurements <strong>of</strong> <strong>the</strong> <strong>in</strong>-plane response<br />

<strong>of</strong> <strong>the</strong> x =0.125 LSCO crystal show that <strong>the</strong> superfluid density is collected from <strong>the</strong> frequency<br />

range below 70–80 cm −1 . This sets up <strong>the</strong> upper bound for <strong>the</strong> magnitude <strong>of</strong> 2∆(0) whereas <strong>the</strong><br />

analysis <strong>of</strong> <strong>the</strong> ω-dependence suggests that gap value can be as small as 40 cm −1 .<br />

[22] Uchida S. et al., Phys. Rev. B, 53 (1996) 14558.<br />

[23] Hosse<strong>in</strong>i A. et al., Phys. Rev. B, 60 (1999) 1349.<br />

[24] Wu D. H. et al., Phys. Rev. Lett., 65 (1990) 2074.<br />

[25] These values are much smaller than typical values measured previously for <strong>the</strong> <strong>vortex</strong> motion<br />

across <strong>the</strong> planes [26], <strong>in</strong> qualitative agreement with a <strong>the</strong>oretical prediction [27] η ‖ /η ⊥ ≈<br />

λ 2 c/3λ 2 ab ∼ 1000–3000. In <strong>the</strong> configuration studied here (see <strong>the</strong> bottom <strong>in</strong>set <strong>of</strong> fig. 2) because<br />

<strong>of</strong> “<strong>in</strong>tr<strong>in</strong>sic p<strong>in</strong>n<strong>in</strong>g” [4], one <strong>in</strong>tuitively expects viscous effects (i.e. dissipation) to be much<br />

smaller than for <strong>the</strong> <strong>vortex</strong> motion across <strong>the</strong> planes.<br />

[26] Golosovsky M. et al., Phys. Rev. B, 50 (1994) 470.<br />

[27] Hao Z. et al., IEEE Trans. Magn., 27 (1991) 1086.<br />

[28] Boeb<strong>in</strong>ger G. S. et al., Phys. Rev. Lett., 77 (1996) 5417.<br />

[29] Arovas D. P. et al., Phys. Rev. Lett., 79 (1997) 2871.<br />

[30] Han J. H. et al., Phys. Rev. Lett., 85 (2000) 1100.<br />

[31] Yang K. and Sondhi S. L., Phys. Rev. B, 57 (1998) 8566.<br />

[32] Wakimoto S. et al., Phys. Rev. B, 62 (2000) 3547; H<strong>of</strong>fman J. E. et al., Science, 295 (2002)<br />

466; Lang K. M. et al., Nature, 415 (2002) 412.<br />

[33] Basov D. N. et al., Science, 283 (1999) 49.<br />

[34] I<strong>of</strong>fe L. B. and Millis A. J., Science, 285 (1999) 1241.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!