Quasiparticle states driven by a scattering on the preformed electron pairs
We analyze evolution of the single particle excitation spectrum of the underdoped cuprate superconductors near the anti-nodal region, considering temperatures below and and above the phase transition. We inspect the phenomenological self-energy that reproduces the angle-resolved-photoemission-spectr...
Збережено в:
| Опубліковано в: : | Condensed Matter Physics |
|---|---|
| Дата: | 2016 |
| Автор: | |
| Формат: | Стаття |
| Мова: | English |
| Опубліковано: |
Інститут фізики конденсованих систем НАН України
2016
|
| Онлайн доступ: | https://nasplib.isofts.kiev.ua/handle/123456789/155774 |
| Теги: |
Додати тег
Немає тегів, Будьте першим, хто поставить тег для цього запису!
|
| Назва журналу: | Digital Library of Periodicals of National Academy of Sciences of Ukraine |
| Цитувати: | Quasiparticle states driven by a scattering on the preformed electron pairs / T. Domanski // Condensed Matter Physics. — 2016. — Т. 19, № 1. — С. 13701: 1–11. — Бібліогр.: 56 назв. — англ. |
Репозитарії
Digital Library of Periodicals of National Academy of Sciences of Ukraine| id |
nasplib_isofts_kiev_ua-123456789-155774 |
|---|---|
| record_format |
dspace |
| spelling |
Domanski, T. 2019-06-17T12:11:40Z 2019-06-17T12:11:40Z 2016 Quasiparticle states driven by a scattering on the preformed electron pairs / T. Domanski // Condensed Matter Physics. — 2016. — Т. 19, № 1. — С. 13701: 1–11. — Бібліогр.: 56 назв. — англ. 1607-324X PACS: 74.20.-z, 74.20.Mn, 74.40.+k DOI:10.5488/CMP.19.13701 arXiv:1601.01592 https://nasplib.isofts.kiev.ua/handle/123456789/155774 We analyze evolution of the single particle excitation spectrum of the underdoped cuprate superconductors near the anti-nodal region, considering temperatures below and and above the phase transition. We inspect the phenomenological self-energy that reproduces the angle-resolved-photoemission-spectroscopy (ARPES) data and we show that above the critical temperature, such procedure implies a transfer of the spectral weight from the Bogoliubov-type quasiparticles towards the in-gap damped states. We also discuss some possible microscopic arguments explaining this process. Проаналiзовано еволюцiю спектру збудження однiєї частинки слабо легованих купратних надпровiдникiв поблизу антинодальної зони, беручи до уваги температури вищi i нижчi за фазовий перехiд. Дослiджено феноменологiчну самоенергiю, яка вiдтворює данi ARPES (кутової фотоемiсiйної спектроскопiї). Показано, що при температурах, вищих за критичну, така процедура передбачає перехiд спектральної ваги вiд квазiчастинок типу Боголюбова до загасаючих станiв всерединi щiлини. Окрiм цього подано певнi мiкроскопiчнi аргументи, якi пояснюють такий процес. Author is indebted for the fruitful discussions with Adam Kamiński, Roman Micnas, Julius Ranninger, and Karol I. Wysokiński. This work is supported by the National Science Centre in Poland through the projects DEC-2014/13/B/ST3/04451. en Інститут фізики конденсованих систем НАН України Condensed Matter Physics Quasiparticle states driven by a scattering on the preformed electron pairs Стани квазiчастинок, керованих розсiюванням на попередньо сформованих електронних парах Article published earlier |
| institution |
Digital Library of Periodicals of National Academy of Sciences of Ukraine |
| collection |
DSpace DC |
| title |
Quasiparticle states driven by a scattering on the preformed electron pairs |
| spellingShingle |
Quasiparticle states driven by a scattering on the preformed electron pairs Domanski, T. |
| title_short |
Quasiparticle states driven by a scattering on the preformed electron pairs |
| title_full |
Quasiparticle states driven by a scattering on the preformed electron pairs |
| title_fullStr |
Quasiparticle states driven by a scattering on the preformed electron pairs |
| title_full_unstemmed |
Quasiparticle states driven by a scattering on the preformed electron pairs |
| title_sort |
quasiparticle states driven by a scattering on the preformed electron pairs |
| author |
Domanski, T. |
| author_facet |
Domanski, T. |
| publishDate |
2016 |
| language |
English |
| container_title |
Condensed Matter Physics |
| publisher |
Інститут фізики конденсованих систем НАН України |
| format |
Article |
| title_alt |
Стани квазiчастинок, керованих розсiюванням на попередньо сформованих електронних парах |
| description |
We analyze evolution of the single particle excitation spectrum of the underdoped cuprate superconductors near the anti-nodal region, considering temperatures below and and above the phase transition. We inspect the phenomenological self-energy that reproduces the angle-resolved-photoemission-spectroscopy (ARPES) data and we show that above the critical temperature, such procedure implies a transfer of the spectral weight from the Bogoliubov-type quasiparticles towards the in-gap damped states. We also discuss some possible microscopic arguments explaining this process.
Проаналiзовано еволюцiю спектру збудження однiєї частинки слабо легованих купратних надпровiдникiв
поблизу антинодальної зони, беручи до уваги температури вищi i нижчi за фазовий перехiд. Дослiджено
феноменологiчну самоенергiю, яка вiдтворює данi ARPES (кутової фотоемiсiйної спектроскопiї). Показано, що при температурах, вищих за критичну, така процедура передбачає перехiд спектральної ваги вiд
квазiчастинок типу Боголюбова до загасаючих станiв всерединi щiлини. Окрiм цього подано певнi мiкроскопiчнi аргументи, якi пояснюють такий процес.
|
| issn |
1607-324X |
| url |
https://nasplib.isofts.kiev.ua/handle/123456789/155774 |
| citation_txt |
Quasiparticle states driven by a scattering on the preformed electron pairs / T. Domanski // Condensed Matter Physics. — 2016. — Т. 19, № 1. — С. 13701: 1–11. — Бібліогр.: 56 назв. — англ. |
| work_keys_str_mv |
AT domanskit quasiparticlestatesdrivenbyascatteringonthepreformedelectronpairs AT domanskit stanikvazičastinokkerovanihrozsiûvannâmnapoperednʹosformovanihelektronnihparah |
| first_indexed |
2025-11-25T21:04:13Z |
| last_indexed |
2025-11-25T21:04:13Z |
| _version_ |
1850543599472607232 |
| fulltext |
Condensed Matter Physics, 2016, Vol. 19, No 1, 13701: 1–11
DOI: 10.5488/CMP.19.13701
http://www.icmp.lviv.ua/journal
Quasiparticle states driven by a scattering
on the preformed electron pairs∗
T. Domański
Institute of Physics, M. Curie-Skłodowska University, 20-031 Lublin, Poland
Received November 13, 2015, in final form November 27, 2015
We analyze evolution of the single particle excitation spectrum of the underdoped cuprate superconductors
near the anti-nodal region, considering temperatures below and and above the phase transition. We inspect the
phenomenological self-energy that reproduces the angle-resolved-photoemission-spectroscopy (ARPES) data
and we show that above the critical temperature, such procedure implies a transfer of the spectral weight
from the Bogoliubov-type quasiparticles towards the in-gap damped states. We also discuss some possible
microscopic arguments explaining this process.
Key words: superconducting fluctutations, Bogoliubov quasiparticles, pseudogap
PACS: 74.20.-z, 74.20.Mn, 74.40.+k
1. Introduction
Superconductivity (i.e., dissipationless motion of the charge carriers) is observed at sufficiently low
temperatures, when electrons from the vicinity of the Fermi surface are bound in the pairs and respond
collectively (rather than individually) to any external perturbation such as electromagnetic field, pres-
sure, temperature gradient, etc. Depending on specific materials, the pairing mechanism can be driven
by phonons (in classical superconductors), magnons (in heavy fermion compounds) or by the antifer-
romagnetic exchange interactions originating from the Coulomb repulsion (in cuprate oxides). In most
cases, the electron pairs are formed at the critical value Tc, marking the onset of superconductivity. There
are, however, numerous exceptions to this rule. For instance, in the cuprate superconductors [1] or in the
ultracold fermionic gasses [2], such pairs pre-exist well above Tc. To some extent, their presence causes
the properties reminiscent of the superconducting state.
Early evidence for the preformed pairs existing above Tc has been indicated in the muon scattering
experiments [3]. Later on, their existence was supported by the ultrafast (tera-Hertz) optical spectroscopy
[4, 5] and the large Nernst effect [6, 7]. Spectroscopic signatures of the preformed pairs have been also
detected directly in the ARPES measurements on yttrium [8] and lanthanum [9] cuprate oxides, revealing
the Bogoliubov-type quasiparticle dispersion above Tc. Furthermore, the STM imaging provided clear
fingerprints of such dispersive Bogoliubov quasiparticles (by the unique octet patterns) being unchanged
from temperatures deep in the superconducting region up to 1.5Tc [10]. Superconducting fluctuations
above Tc have been also reported by the Josephson-like tunneling [11] and the proximity effect induced
in the nanosize metallic slabs deposited on La2−xSrxCuO4 [12]. More recently, the residual Meissner effect
has been experimentally observed above the transition temperature Tc by the torque magnetometry [13]
and other measurements [14, 15]. Additional evidence for the superconducting-like behaviour above Tc
has been seen in the high-resolution ARPES measurements [16], the superfulid fraction observed in the
c-axis optical measurements Re{σc(ω)} [17], the Josephson spectroscopy for YBaCuO-LaSrCuO-YBaCuO
junction using LaSrCuO in the pseudogap state well above Tc [18], optical conductance in the pseudogap
state of YBaCuO superconductor [19] and the photo-enhanced antinodal conductivity in pseudogap state
of the high Tc superconductors [20].
∗This work is dedicated to professor Stefan Sokołowski on the occasion of his 65-th birthday.
© T. Domański, 2016 13701-1
http://dx.doi.org/10.5488/CMP.19.13701
http://www.icmp.lviv.ua/journal
T. Domański
Preformed pairs are correlated above Tc only on some short temporal τφ and spatial lφ scales [21–
23]. For this reason, the superconducting fluctuations are manifested in very peculiar way [24]. Their
influence on the single particle spectrum is manifested by: a) two Bogoliubov-type branches and b) addi-
tional in-gap states that are over-damped (have a short life-time). Temperature has a strong effect on the
transfer of the spectral weight between these entities. In the underdoped cuprate superconductors, such
a transfer is responsible for filling-in the energy gap [16, 25], instead of closing it (as in the classical su-
perconductors). Some early results concerning superconducting fluctuations have been known for a long
time [26, 27], but they attracted much more interest in the context of cuprate superconductors [26–36]
and ultracold fermion superfluids [37, 38].
In this work, we study qualitative changeover of the single particle electronic spectrum of the under-
doped cuprate oxides for temperatures varying from below Tc (in the superconducting state) to above
Tc , where the preformed pairs are not long-range coherent. In the superconducting state, the usual
Bogoliubov-type quasiparticles are driven by the Bose-Einstein condensate of the (zero-momentum)
Cooper pairs. We find that above Tc, the Bogoliubov quasiparticles are still preserved, but the scatter-
ing processes driven by the finite momentum pairs contribute the in-gap states whose life-time substan-
tially increases with increasing temperature. We discuss this process on the phenomenological as well
as microscopic arguments. Roughly speaking, a feedback of the electron pairs on the unpaired electrons
resembles the long-range translational and orientational order that develops between the amphiphilic
particles in presence of the ions at solid state surfaces studied by S. Sokołowski with coworkers [39].
2. Microscopic formulation of the problem
To account for the coherent/incoherent pairing we consider the Hamiltonian
Ĥ =
∑
k,σ
(
εk −µ
)
ĉ†
kσ
ĉkσ+
1
N
∑
k,k′,q
Vk,k′ (q)ĉ†
k′↑ĉ†
q−k′↓ĉq−k↓ĉk↑ (2.1)
describing themobile electrons of kinetic energy εk (where µ is the chemical potential) interacting via the
two-body potential Vk,k′ (q). We assume a separable form Vk,k′ =−g ηk ηk′ of this pairing potential (with
g > 0). In the nearly two-dimensional cuprate superconductors with the prefactor
ηk = 1
2
[
cos(akx )+cos(aky )
]
(where a is the unit length in CuO2 planar structure), such pairing potential
induces the d -wave symmetry order parameter [40, 41].
The Hamiltonian (2.1) can be recast in a more compact form, by introducing the pair operators
b̂q = 1
p
N
∑
k
ηk ĉq−k↓ĉk↑ (2.2)
and b̂†
q = (b̂q)†, when the two-body interactions simplify to
1
N
∑
k,k′,q
Vk,k′(q)ĉ†
k↑ĉ†
q−k↓ĉq−k′↓ĉk′↑ =−
∑
q
g b̂†
qb̂q . (2.3)
Using the Heisenberg equation of motion (ħ= 1)
i
d
dt
ĉk↑ =
(
εk −µ
)
ĉk↑− g ηk
1
p
N
∑
q
ĉ†
q−k↓b̂q (2.4)
we immediately notice that the single-particle properties of this model (2.1):
a) are characterized by the mixed particle and hole degrees of freedom (because the annihilation
operators ĉk↑ couple to the creation operators ĉ†
q−k↓),
b) depend on the pairing field b̂q (appearing in the equation of motion d
dt ĉk↑).
Both these features manifest themselves in the superconducting state, when there exists the Bose-Einstein
(BE) condensate 〈b̂q=0〉 , 0 of the Cooper pairs. They also survive in the normal state, as long as the
preformed (finite-momentum) pairs are present below the same characteristic temperature T ∗ marking
an onset of the electron pairing. In the next section we explore their role in the superconducting state
(T É Tc) and in the pseudogap region (Tc < T < T ∗).
13701-2
Quasiparticle states due to preformed pairs
3. Pairs as the scattering centers
The Heisenberg equation of motion (2.4) indicates that the electronic states are affected by the pairing
field b̂q. Let us consider the generic consequences of such Andreev-type scattering, separately consider-
ing: the BE condensed q = 0 and the finite-momentum q, 0 pairs.
3.1. The effect of the Bose-Einstein condensed pairs
We start by considering the usual BCS approach, when only the zero momentum pairs are taken into
account. This situation has a particularly clear interpretation within the path integral formalism, treating
the pairing field via the Hubbard-Stratonovich transformation and determining it from the minimization
of action (the saddle point solution). The same result can be obtained using the equation of motion (2.4),
focusing on the effect of q = 0 pairs
i
d
dt
ĉk↑ ≃
(
εk −µ
)
ĉk↑− g ηk ĉ†
−k↓
b̂0p
N
, (3.1)
i
d
dt
ĉ†
k↓ ≃ −
(
εk −µ
)
ĉ†
k↓− g ηk
b̂†
0p
N
ĉk↑. (3.2)
Macroscopic occupancy of the q = 0 state implies that the bosonic operators b̂(†)
0 can be treated as complex
numbers b(∗)
0 . By introducing the order parameter
∆k = gηk
〈b̂0〉p
N
= gηk
1
N
∑
k′
ηk′〈ĉ−k′↓ĉk′↑〉 (3.3)
the equations (3.1), (3.2) can be solved exactly using the standard Bogoliubov-Valatin transformation. In
such BCS approach, the classical superconductivity has close analogy with the superfluidity of weakly
interacting bosons, whose collective sound-like mode originates from the interaction between the finite-
momentum bosons and the BE condensate.
In the present context, the zero-momentum Copper pairs substantially affect the single particle ex-
citation spectrum (and the two-body correlations as well). The single-particle Green’s function G(k,τ) =
−T̂τ〈ĉk↑(τ)ĉ†
k↑〉, where T̂τ is the time ordering operator, obeys the Dyson equation
[G(k,ω)]
−1 =ω−εk +µ−Σ(k,ω), (3.4)
with the BCS self-energy
Σ(k,ω) =
|∆k|2
ω+ (εk −µ)
. (3.5)
The self-energy (3.5), accounting for the Andreev-type scattering of the k-momentum electrons on the
Cooper pairs, can be alternatively obtained from the bubble diagram. The related spectral function
A(k,ω) =−π−1ImG(k,ω+ i0+) is thus characterized by the two-pole structure
A(k,ω) = u2
k δ(ω−Ek)+ v2
k δ(ω+Ek) (3.6)
with the Bogoliubov-type quasiparticle energies Ek = ±
√
(
εk −µ
)2 +∆
2
k
and the spectral weights u2
k
=
1
2
[
1+ (εk −µ)/Ek
]
and v2
k
= 1− u2
k
. Let us remark that these quasiparticle branches are separated by
the (true) energy gap |∆k|. In classical superconductors, ∆k implies the off-diagonal-long-range-order
(ODLRO) that quantitatively depends on concentration of the BE condensed Cooper pairs. ODLRO is re-
sponsible for a dissipationless motion of the charge carriers and simultaneously causes the Meissner
effect via the spontaneous gauge symmetry breaking.
13701-3
T. Domański
3.2. The effect of the non-condensed pairs
In this section we shall study effect of the finite-momentum pairs existing above Tc, which no longer
develop any ODLRO because there is no BE condensate. Nevertheless, according to (2.4), the single and
paired fermions are still mutually dependent. This fact suggests that the previous BCS form (3.5) should
be replaced by some corrections originating from the finite momentum pairs. Let us denote the pair
propagator by L(q,τ) =− T̂τ〈b̂(r,τ)b̂†(0,0)〉 and assume its Fourier transform in the following form
L(q,ω) = 1
ω−Eq − i Γ(q,ω)
, (3.7)
where Eq stands for the effective dispersion of pairs and Γ(q,ω) describes the inverse life-time. Taking
into account the equation (2.4), we express the self-energy via the bubble diagram
Σ(k,ω) =−T
∑
iνn ,q
1
ω−ξq−k − iνn
L(q, iνn ), (3.8)
where ξq−k = εq−k −µ and iνn is the bosonic Matsubara frequency. Since above Tc the preformed pairs
are only short-range correlated [21–23], we impose
〈L̂†
(r, t)L̂(0,0)〉∝ exp
(
−
|t |
τφ
−
|r|
ξφ
)
. (3.9)
Following T. Senthil and P.A. Lee [22, 23], one can estimate the single particle Green’s function G (k,ω)
using the following interpolation
Σ(k,ω) =∆
2 ω−ξk
ω2 −
(
ξ2
k
+πΓ2
) , (3.10)
where ∆ is the energy gap due to pairing and the other parameter Γ is related to damping of the subgap
states. In the low energy limit (i.e., for |ω| ≪∆) the dominant contribution comes from the in-gap quasi-
particle whose residue is Z ≡
[
1+∆
2/(πΓ2)
]−1
, whereas at higher energies the BCS-type quasiparticles
are recovered. This selfenergy (3.10) can be derived from the microscopic considerations [42] within the
two-component model, describing itinerant fermions coupled to the hard-core bosons [43–50].
The other (closely relative) phenomenological ansatz [31, 32]
Σ(k,ω) = ∆
2
ω+ξk + iΓ0
− iΓ1 (3.11)
has been inferred considering the “small fluctuations” regime [26]. Experimental lineshapes of the angle
resolved photoemission spectroscopy obtained for the cuprate superconductors at various doping levels
and temperatures (including the pseudogap regime) amazingly well coincide with this simple formula
(3.11). The gap and the phenomenological parameters Γ0, Γ1 are in general momentum-dependent, but
for a given direction in the Brillouin zone one can restrict only to their temperature and doping varia-
tions. From now onwards we shall focus on such antinodal region.
In the overdoped samples, Γ0 can be practically discarded from (3.11) and the remaining parameter
Γ1 simply accounts for T -dependent broadening of the Bogoliubov peaks until they disappear just above
Tc. Physical origin of Γ1 is hence related to the particle-particle scattering. On the contrary, in the under-
doped regime, there exists a pseudogap up to temperatures T ∗ which by far exceed Tc. To reproduce the
experimental lineshapes, one must then incorporate the other parameter Γ0 (nonvanishing only above
Tc) which is scaled by T −Tc as shown in figure 1 reproduced from references [31, 32]. Since Γ0 enters
the self-energy (3.11) through the BCS-type structure, its origin is related to the particle-hole scatterings.
We now inspect some consequences of the parametrization (3.11) applicable for the pseudogap regime
T > Tc in the underdoped cuprates. Since neither the magnitude of Γ1 nor ∆ seem to vary over a large
temperature region above Tc, it is obvious that the qualitative changes are there dominated by scatter-
ings in the particle-hole channel, i.e., due to Γ0. Roughly speaking, these processes are responsible for
filling-in the low energy states upon increasing T as has been evidenced by ARPES [16] and STM [25]
13701-4
Quasiparticle states due to preformed pairs
0
50
100
150
200
0 50 K 100 K 150 K
Temperature
(m
eV
)
∆
Γ0
Γ1
Tc
Figure 1. (Color online) Temperature dependence of the phenomenological parameters ∆, Γ0 and Γ1
which, through the self-energy (3.11), reproduce the experimental profiles of the underdoped Bi2212
(Tc = 83 K) sample. This fitting is adopted from references [31, 32].
measurements. On a microscopic level, such changes can be assigned to scattering on the preformed
pairs.
For analytical considerations, let us rewrite the complex self-energy (3.11) as
Σ(k,ω) = (ω+ξk)
∆
2
(ω+ξk)
2 +Γ
2
0
− iΓk(ω), (3.12)
where the imaginary part is
Γk(ω) = Γ1 +Γ0
∆
2
(ω+ξk)
2 +Γ
2
0
. (3.13)
In what follows we indicate that above Tc the excitation spectrum can consist of altogether three different
states, two of them corresponding to the Bogoliubov modes (signifying particle-hole mixing characteris-
tic for the superconducting state) and another one corresponding to the single particle fermion states
which form inside the pseudogap. These states start to appear at T = T +
c and initially represent heavily
overdamped modes containing infinitesimal spectral weight (see reference [22, 23] for a more detailed
discussion). Upon increasing temperature, their life-time gradually increases and simultaneously the in-
gap states absorb more and more spectral weight at the expense of the Bogoliubov quasiparticles. Finally
(in the particular case considered here, this happens roughly near 2Tc) the single particle fermion states
become dominant.
Anticipating the relevance of (3.11) to the strongly correlated cuprate materials, one can determine
the single particle Green’s function G(k,ω) and the corresponding spectral function A(k,ω). Quasiparticle
energies are determined by poles of the Green’s function, i.e.,
ω−ξk −Re{Σ(k,ω)} = 0 (3.14)
provided that the imaginary part Γk(ω) disappears. We clearly see that the latter requirement cannot be
satisfied for Γ1 , 0 regardless of Γ0. Formally this means that the life-time of herein discussed quasipar-
ticles is not infinite. Let us check these eventual (finite life-time) quasiparticle states determined through
(3.14). Using the self-energy (3.11), the condition (3.14) is equivalent to
(ω−ξk)− (ω+ξk)
∆
2
(ω+ξk)2 +Γ
2
0
= 0. (3.15)
In general, there appear three solutions (figure 2) depending on temperature via the parameter Γ0.
Superconducting region. The fitting procedure [31, 32] has estimated that the parameter Γ0 vanishes
in the superconducting state T É Tc. Under such conditions, (3.15) yields the standard BCS poles at
Ek =±
√
ξ2
k
+∆2. Due to Γ1 , 0, they show up in the spectral function A(k,ω) as Lorentzians whose broad-
ening corresponds to the inverse life-time of the Bogoliubov modes. Owing to T -dependence of Γ1 (see
13701-5
T. Domański
Figure 2. (Color online) Dispersion of ω= Ek representing the solutions of equation (3.14) for T /Tc = 0.95
(dotted line), 1.1, 1.5 and 2 (solid curves as described) obtained for the parameters used in references
[31, 32]. We notice three different crossings (3.15), two of them related to the Bogoliubov modes and
additional one appearing in between.
figure 1), the broadening of these peaks increases upon approaching Tc from below, albeit A(k,ω= 0) = 0.
Experimentally this process can be observed as the smearing of the coherence peaks [1].
Pseudogap regime. With the appearance of Γ0 , 0 above Tc, the real part of the self-energy becomes
a continuous function of ω (see figure 3). Consequently, besides the Bogoliubov modes, we now obtain
an additional crossing located in-between. Figure 2 shows the representative dispersion curves obtained
for 1.1Tc, 1.5Tc, 2Tc and compared with the superconducting state (dotted line). We observe either the
three branches or just the single one at sufficiently high temperatures when the spectral function A(k,ω)
evolves to a single peak structure.
As some useful example, let us study the Fermi momentum kF, when (3.15) simplifies to
ω
(
1− ∆
2
ω2 +Γ
2
0
)
= 0. (3.16)
In this case, we obtain: a) two symmetric quasiparticle energies at ω± =±∆̃, where ∆̃≡
√
∆2 −Γ
2
0
, and b)
the in-gap state at ω0 = 0. The corresponding imaginary parts Γk(ω) are
Γk(ω±) = Γ1 +Γ0 , (3.17)
Γk(ω0) = Γ1 +
∆
2
Γ0
. (3.18)
Figure 3. (Color online) The real part of the self-energy Σ(k,ω) for εk = µ at several representative tem-
peratures T /Tc=0.95 (dotted line) and 1.1, 1.5, 2.0 (as denoted). Below Tc there exist two poles at ω=±∆
whereas for T > Tc, we obtain altogether three crossings which at higher temperaturemerge into a single
one.
13701-6
Quasiparticle states due to preformed pairs
Figure 4. (Color online) The imaginary part Γk(ω) for the same set of parameters as used in figure 3. The
filled symbols indicate the value of Γk(ω) and position of the crossings ω = Ek of the lower Bogoliubov
mode (squares), in-gap state (circles) and the upper Bogoliubov branch (triangles).
Since Γ1 does not vary above Tc, the temperature dependence of Γ−1
k
(ωi ) is controlled by Γ0. Using the ex-
perimental estimations [31, 32], we thus find the qualitatively opposite temperature variations of Γk(ω±)
and Γk(ω0) shown in figure 5. These quantities correspond to the life-times of quasiparticles and, there-
fore, we conclude that:
a) in-gap quasiparticles are forbidden for the superconducting state due to vanishing Γ
−1
k
(ω0) = 0 (in
other words, spectrum consists of just the Bogoliubov modes typical of the BCS theories),
b) in the pseudogap state above Tc, where ∆ , 0 and Γ0 , 0, besides the Bogoliubov branches there
emerge in-gap states which initially at T +
c represent the heavily overdamped modes.
At a first glance, our conclusions seem to be in conflict with the ARPES data, which have not reported
any pronounced in-gap features. Nevertheless, various studies of the pseudogap clearly revealed a rather
negligible temperature dependence of ∆(T ) upon passing Tc (at least for the anti-nodal areas). Instead of
closing this gap, the low energy states are gradually filled-in [25]. Such a behavior can be thought as an
indirect signature of the in-gap states, which for increasing temperatures absorb more and more spectral
weight. To support this conjecture, we illustrate in figure 6 an ongoing transfer of the spectral weights
between the Bogoliubov quasiparticles and in-gap states. Using (3.11), we show the spectral function
A(k,ω) subtracting its value at Tc in analogy to the detailed experimental discussion by T. Kondo et al.
[16]. In-gap states emerge around ω0 and gradually gain the spectral weight (figure 7) simultaneously
increasing their life-time.
Intrinsic broadening of the in-gap states [53] unfortunately obscures their observation by the spectro-
scopic tools at temperatures close to Tc. These states might be, however, probed indirectly. T. Senthil and
P.A. Lee [22, 23] suggested that such states could be responsible for the magnetooscillations observed ex-
0
0.005
0.01
0.015
0.02
0 50 K 100 K 150 K 200 K
Temperature
Γ k
-1
(
m
eV
)-1
In-gap states
Bogoliubov peaks
εk = µ
Figure 5. (Color online) Temperature dependence of the inverse broadening Γ
−1
k
which corresponds to
the effective life-time of the Bogoliubov modes (thin line) and the in-gap states (thick curve) obtained for
εk =µ.
13701-7
T. Domański
-200 -100 0 100 200
ω (meV)
εk = µ
A
(k
,ω
)
-
A
(k
,ω
) |
T
=
T
c
T = 1.1 Tc
T = 1.5 Tc
T = 2.0 Tc
T = 2.5 Tc
Figure 6. (Color online) Transfer of the spectral weight from the Bogoliubov quasiparticle peaks towards
the in-gap states obtained using (3.11) for εk = µ. Temperature dependence of the total transferred spec-
tral weight is shown in figure 7.
perimentally by N. Doiron-Leyraud et al. [51, 52]. They indicated that pair-coherence extending only over
short spatial- and temporal length naturally implies the pair decay (scattering) into the in-gap fermion
states. This line of reasoning has been also followed by some other groups [53, 54].
4. Microscopic toy model
Pairing of the cuprate superconductors occurs on a local scale, practically between the nearest neigh-
bor lattice sites. To account for an interplay between the paired and unpaired charge carriers taking place
in the pseudogap regime we consider here the following simplified picture
Ĥloc = ε0
∑
σ
ĉ†
σĉσ+E0b̂†b̂ +
(
∆b̂†ĉ↓ĉ↑+h.c.
)
, (4.1)
where ĉ(†)
σ correspond to the unpaired fermions and b̂(†) to the pairs (hard-core bosons). We assume that
in the pseudogap state, neither the fermions nor the hard-core boson pairs are long-living because of
their mutual scattering by the Andreev charge exchange term. The same type of scattering, although in
0
1
2
3
4
0.1 1 10
In
-g
ap
s
ta
te
s
(
%
)
T / Tc
Result from eqn (3.11)
Andreev scattering
Figure 7. (Color online) Spectral weight corresponding to the in-gap states obtained from the phenomeno-
logical ansatz (3.11) for εk =µ (dotted curve) and solution of the toymodel (4.1) for ε0 = 0 = E0 (solid line).
13701-8
Quasiparticle states due to preformed pairs
the momentum space, has been considered in reference [22, 23] within the lowest order diagrammatic
treatment. On a microscopic footing, the Hamiltonian (4.1) can be regarded as the effective low energy
description of the plaquettized Hubbard model [47, 48].
Neglecting the itinerancy of the charge carriers, we can obtain a rigorous solution for a given local
cluster (not to be confused with the individual copper sites in CuO2 planes [50]). Exact diagonalization of
the Hilbert space yields the following single particle Green’s function [42]
G(ω) =
ZQP
ω−ε0
+
1−ZQP
ω−ε0 − |∆|2
ω+ε0−E0
. (4.2)
Let us notice that the second term on rhs of (4.2) acquires the same structure as imposed by (3.11). In the
present case, no imaginary terms appear but the structure of the Green’s function (4.2) mimics the role of
Γ0. Formally, it describes the bonding and antibonding states originating from the Andreev scattering and
besides that we also have a remnant of the non-interacting propagator [ω−ε0]−1 whose spectral weight
is given by ZQP.
The quasiparticle weight ZQP depends on occupancies of the fermion and boson levels. As an ex-
ample, we explore here the symmetric (i.e., half-filled) case with ε0 = 0 and E0 = 0 when ZQP = 2/[3+
cosh(|∆|/kBT )] . Assuming the typical ratio |∆|/kBTc = 4, we plot in figure 7 the temperature dependence
of the unpaired states contribution ZQP to the spectrum. We find a very good agreement between our
simple treatment and the estimations using the self-energy (3.11). It means that the parameter Γ0 intro-
duced in references [31, 32] and the local Andreev-type scattering considered here account for the very
same particle-hole processes inducing the in-gap states. Transfer of the spectral weight from the paired
to unpaired states (figure 7) confirms the qualitative agreement over a broad temperature region and the
indication for the same critical point.
For more realistic comparison of the present study with the experimental data [16], one obviously has
to consider the itinerant charge carriers. As a natural improvement of the local solution (4.2) we would
expect the following type of Green’s function
G(k,ω) =
ZQP(k)
ω−εk
+
∑
q
[
1−ZQP(k)
]
f (q,k)
ω−εk − |∆k |2
ω+εq−k−Eq
, (4.3)
where T -dependent coefficients f (q,k) should be determined via the many-body techniques. Approach-
ing Tc from above the predominant influence comes from q → 0 bosons and then we notice that (4.3)
reduces to the ansatz (3.11). Such results have been recently reported from the dynamical mean field
calculations for the Hubbard model [55, 56].
5. Conclusions and outlook
We have shown that the pairing ansatz (3.11), widely used for fitting the experimental ARPES profiles,
above Tc corresponds to the pair scattering inducing the single particle fermion states inside the pseu-
dogap. Temperature dependent phenomenological parameter Γ0 is found to control the transfer of the
spectral weight from the Bogoliubov quasiparticles to the unpaired in-gap states. To model such a pro-
cess on a microscopic level, we have considered the scenario in which the local pairs are scattered into
single fermions via the Andreev conversion [22, 23, 47, 48, 50]. We have found a unique relation between
the transferred spectral weight (from the paired to unpaired quasiparticles) with the non-bonding state
ZQP. It would be instructive to extend the present analysis onto the case of k-dependent energy gap. Such
a problem would be closely related to the issue of Fermi arcs, i.e., partially reconstructed pieces of the
Fermi surface, and to nontrivial angular dependence of the pseudogap [1].
Acknowledgements
Author is indebted for the fruitful discussions with Adam Kamiński, RomanMicnas, Julius Ranninger,
and Karol I. Wysokiński. This work is supported by the National Science Centre in Poland through the
projects DEC-2014/13/B/ST3/04451.
13701-9
T. Domański
References
1. Kaminski A., Kondo T., Takeuchi T., Gu G., Philos. Mag. B, 2014, 95, 453; doi:10.1080/14786435.2014.906758.
2. Gaebler J.P., Stewart J.T., Drake T.E., Jin D.S., Perali A., Pieri P., Strinati G.C., Nat. Phys., 2010, 6, 569;
doi:10.1038/nphys1709.
3. Uemura Y.J., Le L.P., Luke G.M., Sternlieb B.J., Wu W.D., Brewer J.H., Riseman T.M., Seaman C.L., Maple M.B.,
Ishikawa M., Hinks D.G., Jorgensen J.D., Saito G., Yamochi H., Phys. Rev. Lett., 1991, 66, 2665;
doi:10.1103/PhysRevLett.66.2665.
4. Orenstein J., Corson J., Oh S., Eckstein J.N., Ann. Phys. (Leipzig), 2006, 15, 596; doi:10.1002/andp.200510202.
5. Corson J., Mallozzi R., Orenstein J., Eckstein J.N., Bozovic I., Nature, 1999, 398, 221; doi:10.1038/18402.
6. Wang Y., Li L., Ong N.P., Phys. Rev. B, 2006, 73, 024510; doi:10.1103/PhysRevB.73.024510.
7. Xu Z.A., Ong N.P., Wang Y., Takeshita T., Uchida S., Nature, 2000, 406, 486; doi:10.1038/35020016.
8. Kanigel A., Chatterjee U., Randeria M., NormanM.R., Koren G., Kadowaki K., Campuzano J., Phys. Rev. Lett., 2008,
101, 137002; doi:10.1103/PhysRevLett.101.137002.
9. Shi M., Bendounan A., Razzoli E., Rosenkranz S., Norman M.R., Campuzano J.C., Chang J., Mansson M., Sassa Y.,
Claesson T., Tjernberg O., Patthey L., Momono N., Oda M., Ido M., Guerrero S., Mudry C., Mesot J., Eur. Phys. Lett.,
2009, 88, 27008; doi:10.1209/0295-5075/88/27008.
10. Lee J., Fujita K., Schmit A.R., Kim C.K., Eisaki H., Uchida S., Davis J.C., Science, 2009, 325, 1099;
doi:10.1126/science.1176369.
11. Bergeal N., Lesueur J., Aprili M., Faini G., Contour J.P., Leridon B., Nat. Phys., 2008, 4, 608; doi:10.1038/nphys1017.
12. Yuli O., Asulin I., Kalchaim Y., Koren G., Millo O., Phys. Rev. Lett., 2009, 103, 197003;
doi:10.1103/PhysRevLett.103.197003.
13. Li L., Wang Y., Komiya S., Ono S., Ando Y., Gu G.D., Ong N.P., Phys. Rev. B, 2010, 81, 054510;
doi:10.1103/PhysRevB.81.054510.
14. Iye T., Nagatochi T., Ikeda R., Matsuda A., J. Phys. Soc. Jpn., 2010, 79, 114711; doi:10.1143/JPSJ.79.114711.
15. Bernardi E., Lascilfari A., Ragimonti A., Romano L., Scavini M., Oliva C., Phys. Rev. B, 2010, 81, 064502;
doi:10.1103/PhysRevB.81.064502.
16. Kondo T., Khasanov R., Takeuchi T., Schmalian J., Kaminski A., Nature, 2009, 457, 296; doi:10.1038/nature07644.
17. Dubroka A., Rössle M., Kim K.W., Malik V.K., Munzar D., Basov D.N., Schafgans A.A., Moon S.J., Lin C.T.,
Haug D., Hinkov V., Keimer B., Wolf Th., Storey J.G., Tallon J.L., Bernhard C., Phys. Rev. Lett., 2011, 106, 047006;
doi:10.1103/PhysRevLett.106.047006.
18. Kirzhner T., Koren G., Sci. Rep., 2014, 4, 6244; doi:10.1038/srep06244.
19. Moon S.J., Lee Y.S., Schafgans A.A., Chubukov A.V., Kasahara S., Shibauchi T., Terashima T., Matsuda Y.,
Tanatar M.A., Prozorov R., Thaler A., Canfield P.C., Bud’ko S.L., Sefat A.S., Mandrus D., Segawa K., Ando Y.,
Basov D.N., Phys. Rev. B, 2014, 90, 014503; doi:10.1103/PhysRevB.90.014503.
20. Cilento F., Dal Conte S., Coslovich G., Peli S., Nembrini N., Mor S., Banfi F., Ferrini G., Eisaki H., Chan M.K.,
Dorow C.J., Veit M.J., Greven M., van der Marel D., Comin R., Damascelli A., Rettig L., Bovensiepen U., Capone M.,
Giannetti C., Parmigiani F., Nat. Commun., 2014, 5, 4353; doi:10.1038/ncomms5353.
21. Franz M., Nat. Phys., 2007, 3, 686; doi:10.1038/nphys739.
22. Senthil T., Lee P.A., Phys. Rev. Lett., 2009, 103, 076402; doi:10.1103/PhysRevLett.103.076402.
23. Senthil T., Lee P.A., Phys. Rev. B, 2009, 79, 245116; doi:10.1103/PhysRevB.79.245116.
24. Mamedov T.A., de Llano M., Philos. Mag. B, 2014, 94, 4102; doi:10.1080/14786435.2014.979903.
25. Fischer O., Kugler M., Maggio-Aprile I., Berthod Ch., Renner Ch., Rev. Mod. Phys., 2007, 79, 353;
doi:10.1103/RevModPhys.79.353.
26. Abrahams E., Red M., Woo J.W.F., Phys. Rev. B, 1970, 1, 208; doi:10.1103/PhysRevB.1.208.
27. Schmid A., Z. Phys., 1970, 231, 324; doi:10.1007/BF01397514.
28. Ranninger J., Robin J.M., Physica C, 1995, 253, 279; doi:10.1016/0921-4534(95)00515-3.
29. Emery V.J., Kivelson S.A., Nature, 1995, 374, 434; doi:10.1038/374434a0.
30. Tchernyshyov O., Phys. Rev. B, 1997, 56, 3372; doi:10.1103/PhysRevB.56.3372.
31. Norman M.R., Randeria M., Ding H., Campuzano J.C., Phys. Rev. B, 1998, 57, 11093;
doi:10.1103/PhysRevB.57.R11093.
32. Franz M., Millis A.J., Phys. Rev. B, 1998, 58, 14572; doi:10.1103/PhysRevB.58.14572.
33. Fujimoto S., J. Phys. Soc. Jpn., 2002, 71, 1230; doi:10.1143/JPSJ.71.1230.
34. Domański T., Ranninger J., Phys. Rev. Lett., 2003, 91, 255301; doi:10.1103/PhysRevLett.91.255301.
35. Domański T., Ranninger J., Phys. Rev. B, 2001, 63, 134505; doi:10.1103/PhysRevB.63.134505.
36. Chubukov A.V., Norman M.R., Millis A.J., Abrahams E., Phys. Rev. B, 2007, 76, 180501;
doi:10.1103/PhysRevB.76.180501.
37. Chen Q.J., Stajic J., Tan S.N., Levin K., Phys. Rep., 2005, 412, 1; doi:10.1016/j.physrep.2005.02.005.
13701-10
http://dx.doi.org/10.1080/14786435.2014.906758
http://dx.doi.org/10.1038/nphys1709
http://dx.doi.org/10.1103/PhysRevLett.66.2665
http://dx.doi.org/10.1002/andp.200510202
http://dx.doi.org/10.1038/18402
http://dx.doi.org/10.1103/PhysRevB.73.024510
http://dx.doi.org/10.1038/35020016
http://dx.doi.org/10.1103/PhysRevLett.101.137002
http://dx.doi.org/10.1209/0295-5075/88/27008
http://dx.doi.org/10.1126/science.1176369
http://dx.doi.org/10.1038/nphys1017
http://dx.doi.org/10.1103/PhysRevLett.103.197003
http://dx.doi.org/10.1103/PhysRevB.81.054510
http://dx.doi.org/10.1143/JPSJ.79.114711
http://dx.doi.org/10.1103/PhysRevB.81.064502
http://dx.doi.org/10.1038/nature07644
http://dx.doi.org/10.1103/PhysRevLett.106.047006
http://dx.doi.org/10.1038/srep06244
http://dx.doi.org/10.1103/PhysRevB.90.014503
http://dx.doi.org/10.1038/ncomms5353
http://dx.doi.org/10.1038/nphys739
http://dx.doi.org/10.1103/PhysRevLett.103.076402
http://dx.doi.org/10.1103/PhysRevB.79.245116
http://dx.doi.org/10.1080/14786435.2014.979903
http://dx.doi.org/10.1103/RevModPhys.79.353
http://dx.doi.org/10.1103/PhysRevB.1.208
http://dx.doi.org/10.1007/BF01397514
http://dx.doi.org/10.1016/0921-4534(95)00515-3
http://dx.doi.org/10.1038/374434a0
http://dx.doi.org/10.1103/PhysRevB.56.3372
http://dx.doi.org/10.1103/PhysRevB.57.R11093
http://dx.doi.org/10.1103/PhysRevB.58.14572
http://dx.doi.org/10.1143/JPSJ.71.1230
http://dx.doi.org/10.1103/PhysRevLett.91.255301
http://dx.doi.org/10.1103/PhysRevB.63.134505
http://dx.doi.org/10.1103/PhysRevB.76.180501
http://dx.doi.org/10.1016/j.physrep.2005.02.005
Quasiparticle states due to preformed pairs
38. Cichy A., Micnas R., Ann. Phys. (New York), 2014, 347, 207; doi:10.1016/j.aop.2014.04.014.
39. Pizio O., Sokołowski S., Sokołowska Z., J. Chem. Phys., 2014, 140, 174706; doi:10.1063/1.4873438.
40. Matsui H., Sato T., Takahashi T., Wang S.-C., Yang H.-B., Ding H., Fujii T., Watanabe T., Matsuda A., Phys. Rev. Lett.,
2003, 90, 217002; doi:10.1103/PhysRevLett.90.217002.
41. Chatterjee U., Shi M., Ai D., Zhao J., Kanigel A., Rosenkranz S., Raffy H., Li Z.Z., Kadowaki K., Hinks D.G.,
Xu Z.J., Wen J.S., Gu G., Lin C.T., Claus H., Norman M.R., Randeria M., Campuzano J., Nat. Phys., 2009, 5, 1456;
doi:10.1038/nphys1200.
42. Domański T., Phys. Rev. A, 2011, 84, 023634; doi:10.1103/PhysRevA.84.023634.
43. Friedberg R., Lee T.D., Phys. Rev. B, 1989, 40, 6740; doi:10.1103/PhysRevB.40.6745.
44. Micnas R., Ranninger J., Robaszkiewicz S., Rev. Mod. Phys., 1990, 62, 113; doi:10.1103/RevModPhys.62.113.
45. Micnas R., Phys. Rev. B, 2007, 76, 184507; doi:10.1103/PhysRevB.76.184507.
46. Geshkenbein V.B., Ioffe L.B., Larkin A.I., Phys. Rev. B, 1997, 55, 3173; doi:10.1103/PhysRevB.55.3173.
47. Altman E., Auerbach A., Phys. Rev. B, 2002, 65, 104508; doi:10.1103/PhysRevB.65.104508.
48. Mihlin A., Auerbach A., Phys. Rev. B, 2009, 80, 134521; doi:10.1103/PhysRevB.80.134521.
49. Le Hur K., Rice T.M., Ann. Phys. (New York), 2009, 324, 1452; doi:10.1016/j.aop.2009.02.004.
50. Ranninger J., Domański T., Phys. Rev. B, 2010, 81, 014514; doi:10.1103/PhysRevB.81.014514.
51. Doiron-Leyraud N., Proust C., LeBoeuf D., Levallais J., Bonnemaison J.-B., Liang R., Bonn D.A., Hardy W.N., Taille-
fer L., Nature, 2007, 447, 565; doi:10.1038/nature05872.
52. Taillefer L., J. Phys.: Condens. Matter, 2009, 21, 164212; doi:10.1088/0953-8984/21/16/164212.
53. Ranninger J., Romano A., Phys. Rev. B, 2010, 82, 054508; doi:10.1103/PhysRevB.82.054508.
54. Micklitz T., Norman M.R., Phys. Rev. B, 2009, 80, 220513; doi:10.1103/PhysRevB.80.220513.
55. Imada M., Yamaji Y., Sakai S., Motome Y., Ann. Phys. (Berlin), 2001, 523, 629; doi:10.1002/andp.201100028.
56. Sakai S., Civelli M., Nomura Y., Imada M., Phys. Rev. B, 2015, 92, 180503(R); doi:10.1103/PhysRevB.92.180503.
Стани квазiчастинок, керованих розсiюванням
на попередньо сформованих електронних парах
T. Доманьский
Iнститут фiзики, Унiверситет Марiї Кюрi-Складовської, 20-031 Люблiн, Польща
Проаналiзовано еволюцiю спектру збудження однiєї частинки слабо легованих купратних надпровiдникiв
поблизу антинодальної зони, беручи до уваги температури вищi i нижчi за фазовий перехiд. Дослiджено
феноменологiчну самоенергiю, яка вiдтворює данi ARPES (кутової фотоемiсiйної спектроскопiї). Показа-
но, що при температурах, вищих за критичну, така процедура передбачає перехiд спектральної ваги вiд
квазiчастинок типу Боголюбова до загасаючих станiв всерединi щiлини. Окрiм цього подано певнi мiкро-
скопiчнi аргументи, якi пояснюють такий процес.
Ключовi слова: флуктуацiїї надпровiдностi, квазiчастинки Боголюбова, псевдощiлина
13701-11
http://dx.doi.org/10.1016/j.aop.2014.04.014
http://dx.doi.org/10.1063/1.4873438
http://dx.doi.org/10.1103/PhysRevLett.90.217002
http://dx.doi.org/10.1038/nphys1200
http://dx.doi.org/10.1103/PhysRevA.84.023634
http://dx.doi.org/10.1103/PhysRevB.40.6745
http://dx.doi.org/10.1103/RevModPhys.62.113
http://dx.doi.org/10.1103/PhysRevB.76.184507
http://dx.doi.org/10.1103/PhysRevB.55.3173
http://dx.doi.org/10.1103/PhysRevB.65.104508
http://dx.doi.org/10.1103/PhysRevB.80.134521
http://dx.doi.org/10.1016/j.aop.2009.02.004
http://dx.doi.org/10.1103/PhysRevB.81.014514
http://dx.doi.org/10.1038/nature05872
http://dx.doi.org/10.1088/0953-8984/21/16/164212
http://dx.doi.org/10.1103/PhysRevB.82.054508
http://dx.doi.org/10.1103/PhysRevB.80.220513
http://dx.doi.org/10.1002/andp.201100028
http://dx.doi.org/10.1103/PhysRevB.92.180503
Introduction
Microscopic formulation of the problem
Pairs as the scattering centers
The effect of the Bose-Einstein condensed pairs
The effect of the non-condensed pairs
Microscopic toy model
Conclusions and outlook
|