SciPost logo

SciPost Submission Page

Thermodynamics of the thermoelectric working fluid close to the superconducting phase transition

by I. Khomchenko, A. Ryzhov, F. Maculewicz, F. Kurth, R. Hühne, A. Golombek, M. Schleberger, C. Goupil, Ph. Lecoeur, G. Benenti, G. Schierning, H. Ouerdane

This is not the latest submitted version.

Submission summary

Authors (as registered SciPost users): Giuliano Benenti · Ilia Khomchenko · Henni Ouerdane · Gabi Schierning
Submission information
Preprint Link: https://arxiv.org/abs/2110.11000v2  (pdf)
Date submitted: 2022-04-14 07:54
Submitted by: Ouerdane, Henni
Submitted to: SciPost Physics
Ontological classification
Academic field: Physics
Specialties:
  • Condensed Matter Physics - Experiment
  • Condensed Matter Physics - Theory
Approaches: Theoretical, Experimental

Abstract

The bottleneck in state-of-the-art thermoelectric power generation and cooling is the low performance of thermoelectric materials. While the adverse effects of lattice phonons on performance can be mitigated, the main difficulty remains to obtain a large thermoelectric power factor as the Seebeck coefficient and the electrical conductivity cannot be increased independently. Here, relating the thermoelastic properties of the electron gas that performs the thermoelectric energy conversion, to its transport properties, we analyze theoretically whether an electronic phase transition can enhance thermoelectric conversion and at what cost. More precisely, we consider the metal-to-superconductor phase transition in a model two-dimensional system, and we seek to quantify the contribution of the 2D fluctuating Cooper pairs to the power factor in the close vicinity of the critical temperature $T_{\rm c}$. In addition, we provide experimental evidence of the rapid increase of the Seebeck coefficient without decreasing the electrical conductivity near $T_{\rm c}$ in a 100-nm Ba(Fe$_{1-x}$Co$_x$)$_2$As$_2$ thin film with high structural quality resulting in a power factor enhancement of approximately 300. This level of performance cannot be achieved in a system with low structural quality as shown experimentally with our sample degraded by ion bombardment as defects preclude the strong enhancement of the Seebeck coefficient near the phase transition. Finally, we theoretically discuss the ideal thermoelectric conversion efficiency (i.e. disregarding adverse phonon effects) and show that driving the electronic system to the vicinity of a phase transition may be an innovative path towards a strong performance increase but at the cost of a narrow temperature range of use of such materials.

Current status:
Has been resubmitted

Reports on this Submission

Anonymous Report 2 on 2022-9-2 (Invited Report)

Strengths

Interesting work. The experimental part is well explained.

Weaknesses

Theory part is badly explained. I have no reason to think there is anything wrong, but certain crucial information is missing (see report). This can be rectified by adding sentences at various points in the manuscript.

Report

Report on Khomchenko et al "Thermodynamics of the thermoelectric working fluid close to the
superconducting phase transition"

This work has an experimental part and a theory part.
I will discuss them separately.

EXPERIMENT:
This is an interesting work which measure the thermoelectric properties of
superconducting thin-film made of Ba(Fe_{1−x} Co_x)_2 As_2 (with 100nm thickness).
Under two conditions
Fig 3a : higher structural quality (lower disorder)
Fig 3b : lower structural quality (higher disorder)

The lower structural quality is created by the ion-bombardment of samples with higher structural quality;
the difference between higher and lower structural quality is about 10%,
That is to say that the ion-bombardment raises the resistivity of the thin-film's normal state at T ~ 30K
(i.e. just above the superconducting transition) by about 10% from about 10 to 11 micro-Ohm-metres.
Intriguingly, this modest change in structural quality has a huge effect on the thermoelectric response;
the 10% reduction of the structural quality HALVES the thermoelectric response!
That is to say that magnitude of the Seebeck coefficient at the superconducting transition changes
from 200 to 100 microvolts/K.

Looking at the Seebeck curve in Fig 3b, it looks like a smoothed version of that in Fig 3a.
Thus I *speculate* that the ion bombardment is causing the superconducting transition temperature, Tc, to vary randomly across the sample. It would be interesting to make a theoretical model in which each square of the sample has the alpha (T) of the shape in Fig 3a, but with Tc varying randomly from square to square.
One could see if this reproduces the curve in Fig 3b.

THEORY:
I found the theory part very hard to follow,
the notation was confusing, and some crucial information seems to be missing.
I list the things that confused me here, in the hope it will help the authors add sentences here and there to make the manuscript easier to understand.

(A) CONFUSION IN DEFINITION OF ZT VERSUS Z_{th}T

Eq. (1) tells us that ZT is the figure of merit including electrons AND phonons, and
paragraph one of section 2.1 that Z_{th}T is that of the electrons alone (neglecting phonons).
However in Eq. (2) Z_{th}T is defined via the isentropic expansion factor,
and ZT becomes the figure of merit of the electrons alone in Eq. (3).
This is very confusing, please fix the notation to be more consistent.
I think one actually needs 3 different symbols for three different "ZT"s;
- figure of merit including electrons AND phonons
- figure of merit for electrons alone
- figure of merit defined via the isentropic expansion factor

(B) MISSING INFORMATION FOR ZT

Unless I missed it, the model of the "2d electron gas" in Fig 4, 5 is incomplete.
Firstly Fig 4 is misleading in implying that only kappa changes between the two plots,
because alpha, sigma, C_{mu} and C_{N} are all also completely different for "2d electron gas" and "FCP".
Then the appendices do not give the reader enough information to know kappa_{e}, alpha_{e}, etc,
because the authors do not give the form of the terms inside Sigma(E) below eq. (16).
The most common assumption for a 2d electron gas model for a thin-film of metal is that the temperature is much less than the Fermi energy, so Sigma is almost energy independent over the range of integration in Eq. (15) (Sommerfeld expansion), making the Seebeck coefficient extremely small.
This does not seem to be the case here, but I cannot guess more without knowing what parameters are assumed.
Specifically;
(i) What is temperature range, Fermi energy, etc for "2d electron gas" plot in Fig 4?
(ii) What is Fermi energy for "2DEG" plot in Fig 5, and "1D", "2D", "3D" plots in Fig. 6?
Even better would be plots of alpha_{e}, kappa{e}, etc, for the authors' chosen parameters.
(iii) What is the model for 0D in Fig 6,7? What is its density of states (DOS)?
Is it single-level system, with a delta-function DOS, or a Lorentzian-DOS due to broadening by coupling to leads?
Or is it a multi-level system?

(C) MISSING INFORMATION FOR Z_{th}T FOR FLUCTUATING COOPER PAIRS (FCP).

Unless I missed it, the information enabling one to calculate Z_{th}T from Eq. (2)
is incomplete for the fluctuating Cooper pairs (FCP).
At least, it is very hard to find the information in the manuscript.
I see that appendix A contains formulas for some of the quantities that enter Eq. (2),
but I did not find a formula for beta or C_{mu}.
Similarly, I did not find a formula for N_{cp},
which enter Xi_{T} in Eq (12), necessary for Eq. (3).
In contrast,
I found the following formula for Z_{th}T for
fluctuating Cooper pairs in Ref [15] (Ouerdane et al PRB, 2015)
Z_{th}= ln[1/epsilon]
Is that what is used here? If so, it would help the reader to give this formula!


(D) PHYSICS OF FLUCTUATING COOPER PAIRS

It would greatly help the reader if the authors included a paragraph that outlined the basic physics
contained in the fluctuating cooper pairs (FCP) theory.

(E) CONDITION FOR NEGLECTING LATTICE PHONONS, kappa_{lat}

I am confused by whether it is reasonable to take kappa=kappa_{cp} rather than
kappa=kappa_{cp}+kappa=kappa_{lat} in the right-hand plot in Fig 4.
We see from Eq. (24) that kappa_{cp} goes to zero at the critical point, and so it will
become less than kappa_{lat} at some point.
Have the authors estimated when this occurs,
and made sure that it is reasonable to take kappa=kappa_{cp} for the temperature range in Fig 4?
(As stated above, I could not find the range of temperature or epsilon for this plot in the manuscript).

Requested changes

See main report

  • validity: good
  • significance: good
  • originality: good
  • clarity: low
  • formatting: excellent
  • grammar: excellent

Author:  Henni Ouerdane  on 2023-07-25  [id 3836]

(in reply to Report 2 on 2022-09-02)
Category:
answer to question
reply to objection

We thank the Referee for an overall positive evaluation. We also take good note of the criticism and the various points raised, which we address in our reply.

EXPERIMENT

We appreciate a very in-depth look into and discussion of our experimental data. Since the manuscript has meanwhile become relatively long, in particular also due to necessary additions by the revision, we have not further deepened these very interesting aspects that can be considered in future works.

THEORY

(A) Confusion between $ZT$ and $Z_{\rm th}T$

Our reply:

The traditional notations used for the thermoelectric figure of merit are usually $ZT$ or $zT$, and they include the thermal conductivity of the lattice in the definition. In our work, as in some others where authors adapt their notations accordingly, we focus solely on the conversion performance of the electron gas. To avoid confusion we adopt the following notations in the revised version of the manuscript:

* $zT$ for the standard thermoelectric figure of merit, given by Eq. (1) of the manuscript;

* $z_{\rm e}T$ for the thermoelectric figure of merit \emph{disregarding} the phonon effects (i.e. without $\kappa_{\rm lat}$). It is a measure of the electron gas performance alone. Hence, the quantity $z_{\rm e}T$ includes only the Seebeck coefficient, the electrical conductivity, and the electronic thermal conductivity; and, of course, $z_{\rm e}T > zT$. In fact, in a given material, $z_{\rm e}T$ is an upper bound for $zT$ that would be reached in a situation where the so-called ``phonon glass'' (absence of heat transport by the lattice) would be managed. It is given by Eq. (2) of the revised manuscript;

* $Z_{\rm th}T$ is what we call the thermodynamic figure of merit defined from the electron gas heat capacity ratio, in Eq. (21). In the present work, we related the thermoelastic properties of the electrons to their transport properties -- see the parametric plots $z_{\rm e}T$ vs $Z_{\rm th}$ in Fig. 1 of the revised manuscript (Fig. 4 of the former version);

* $z_{\rm cp}T$ is the figure of merit of the 2D fluctuating Cooper pairs system, when the system's temperature $T$ is in the close vicinity of $T_{\rm c}$. It is given by Eq. (23);

* $Z_{\rm th,cp}T$ is what we call the thermodynamic figure of merit defined from the 2D FCP system's heat capacity ratio. It is given by Eq. (22).

(B) Missing information for $ZT$

Our reply:

In Fig. 4 of the former version of the manuscript (now Fig. 1), the $ZT$s'' and the$Z_{\rm th}T$s'' are those computed with the transport coefficients and thermoelastic coefficients of the 2D electron gas and the 2D FCP respectively. All the transport and thermoelastic coefficients used for the computation of the electron gas $ZT$'' (now $z_{\rm e}T$ and $Z_{\rm th}T$) are, indeed, completely different from those used to compute the 2D FCP$ZT$'' (now $z_{\rm cp}T$) and $Z_{\rm th}T$ (now $Z_{\rm th,cp}$). The same applies for the power factors in Fig. 5 of the former version of the manuscript (now Fig. 3), though we kept a generic notation as the arrows in the plot clearly indicate to what model the curves correspond.

The transport distribution function $\Sigma(E)$ has three terms: the relaxation time $\tau$, which for simplicity we take as constant, the velocity $v \propto \sqrt{E}$, and the density of states $g(E)$ for noninteracting electron systems. More detail is now given in Appendix A of the revised version.

For our numerical simulations, we consider low-density electron gases with concentrations $n_{\rm 3D}=10^{18} \mathrm{cm}^{-3}$, $n_{\rm 2D}=10^{12} \mathrm{cm}^{-2}$, and $n_{\rm 1D}=10^{6} \mathrm{cm}^{-1}$ for the three-, two-, and one-dimensional systems, respectively.

(i) The temperature ranges from close to $0$ K to $300$ K for the numerical calculations of the electron gases, and from $T_{\rm c}$ to $T_{\rm c}$ + 0.2 K for the 2D FCP.

(ii) The Fermi energy for the two-dimensional electron gas, is $E_{\rm F}^{\rm 2D}=2.39$ meV. The value of the Fermi energy in the three-dimensional case is $E_{\rm F}^{\rm 3D}=3.64$ meV, and $E_{\rm F}^{\rm 1D}=0.94$ meV for the one-dimensional case equals.

(iii) We use a single-level quantum dot to model 0D electron gas and a Lorentzian form for the density of states:

$g_{0D}(E) = \frac{\Gamma}{(E - E_{0})^{2} +(\Gamma/2)^{2}}$

where $\Gamma$ is the energy level width. The central energy of the channel is $E_{0}=0.99E_{F}^{\rm 2D}$, and the channel coupling energy $\Gamma = 0.1k_{B}T$.

(C) Missing information for $Z_{th}T$ for fluctuating Cooper pairs (FCP).

Our reply:

In the revised version, we transferred parts of the appendices of the previous version to the main text. In fact, as this work contains much material that pertains to thermodynamics, materials science, coupled transport, modeling, experiments etc., we wanted, for ease of reading, to avoid an overload of definitions and formulas that can be found in the cited published references.

To answer more specifically the point on the $Z_{\rm th}$ for the 2D FCP, we use and adapt the definition of $\gamma$ in Eq. (21) of the revised version, which explicitly reads:

\[ \gamma_{\rm cp} = 1 + \frac{\alpha_{\rm th,cp}}{\ell_{\rm cp}} = 1 +Z_{\rm th,cp}T \]

\noindent where $\ell_{\rm cp} = C_{N_{\rm cp}}/(q^2\chi_{T_{\rm cp}} T) = -C_{\rm GL} k_{\rm B}^2\ln\epsilon/q^2$, with $C_{\rm GL} = \hbar^2/(2m_{\rm cp}k_{\rm B}T_{\rm c}\xi^2)$ a dimensionless parameter in the Ginzburg-Landau free energy functional. Here $\xi$ is the coherence length. This is how we get to $Z_{\rm th,cp}T = \gamma_{\rm cp} - 1 = -\ln\epsilon$. See also the calculations shown in the appendix of Ref. [15] of the manuscript (Ouerdane et al PRB, 2015).

(D) Physics of fluctuating Cooper pairs

Our reply:

In the new section 2.2.3 of the revised manuscript, we give more information on the nature of the fluctuating Cooper pairs. The physics of fluctuating Cooper pairs is introduced and discussed at length in Ref. [42] of the revised manuscript: A. Larkin and A. Varlamov, Theory of Fluctuations in Superconductors, revised edition, (Oxford Science publications, 2009).

(E) Condition for neglecting lattice phonons $kappa_{lat}$

Our reply:

Interesting point here, which we now discuss in the new Section 2.4. Though lattice phonons are to be considered for a complete evaluation of a thermoelectric material's performance, they are primarily a cause of heat leaks, hence energy loss. It has been stated many times in thermoelectricity papers that the ideal thermoelectric material would behave as an electron crystal–phonon glass''\footnote{This paradigm was first expressed in G. A. Slack, CRC Handbook of Thermoelectrics, edited by D. M. Rowe (CRC Press, Boca Raton, FL, 1995), p. 407.} system, meaning high electrical conductivity and very low, nay zero thermal conductivity. So, more specifically, in the present work where we relate the thermoelastic properties to the transport properties of the electronic working fluid alone in Fig. 4 of the former manuscript (now Fig. 1 in the revised version with the new notations), we evaluate what the maximum conversion efficiency can ideally be in absence of heat leaks. In the case of 2D FCP, yes $\kappa_{\rm cp} \rightarrow 0$ in the limit $T \rightarrow T_{\rm c}$, but this does not pose any problem, quite the contrary in fact, as this shows what happens if the system would tend to become like anelectron crystal'', making $z_{\rm cp}T$ larger. Sure, accounting for $\kappa_{\rm lat}$ would negatively impact on the increase of $z_{\rm cp}T$, but here we see that without heat leaks, the conversion efficiency can tend to Carnot's efficiency in the limit $T\rightarrow T_{\rm c}$. The range over which the numerical calculations are performed in the normal phase is $\left]T_{\rm c} = 25.6 ~;~ 300 \right]$ K, and in the superconducting fluctuation regime $\left]T_{\rm c}~;~ T_{\rm c}+\delta T\right]$, with $\delta T = 0.2$ K.

Anonymous Report 1 on 2022-8-4 (Invited Report)

Report

The manuscript “Thermodynamics of the thermoelectric working fluid close to the superconducting phase transition” by I. Khomchenko et al. is a study of the thermoelectric properties in a metal at the verge of a superconducting phase transition. According to the presented calculations an electronic system in such a state could offer a large enhancement of the thermoelectric energy conversion efficiency. To support their theoretical results, the authors performed measurements of the thermoelectric power in the 100-nm thick Ba(Fe1-xCox)2As2 thin film.
A relatively small efficiency of the currently used thermoelectric power generators drives intensive research to improve their thermal energy conversion ability. Since the figure of merit includes both the Seebeck coefficient (S) and the thermal conductivity, the effort is usually made to increase the former or supress the latter. In the present manuscript the authors chose the first path, concluding that near the superconducting transition S should be enhanced. As it has been already shown, both the transverse and longitudinal thermoelectrical coeffcients are expected to increase in the presence of fluctuating Cooper pairs [I. Ussishkin et al., Phys. Rev. Lett. 89, 287001 (2002)]. This applies to both 2D and 3D case. While the effect was shown to be likely responsible for the enhancement of the Nernst effect in the hight-Tc superconductors, the influence on the Seebeck was not noticed neither in bulk samples [e.g. S.D. Obertelli et al., Phys. Rev. B 46, 14928(R) (1992)] nor in thin films [e.g. H.-C. Ri et al., Phys. Rev. B 50, 3312 (1994), M. Putti et al., Int. J. Mod. Phys. B 17, 415 (2003)] of cuprates. The effect is absent also in the iron-based superconductors [I. Pallecchi et al., Supercond. Sci. Technol. 29, 073002 (2016)]. The authors tried to provide some references to experimental evidence, but “broadened” peaks from Refs. [26, 27, 28, 39, 40] look more like counterexamples. I mean that the Seebeck coefficient should not decrease when approaching critical temperature, if its magnitude were related to fluctuating Cooper pairs. Moreover, judging by the temperature where the absolute value of S start to increase in the given examples, the phenomenon is unlikely related to superconducting fluctuations. Another thing is that the sign of the excessive Seebeck coefficient, which I believe should be positive for a 2D case, is oppositely to that observed in Ba(Fe1-xCox)2As2. On the other hand, there is a good evidence of the enhancement of the thermoelectric power factor in ultra-thin FeSe [S. Shimizu et al., Nature Communications 10, 825 (2019)], but it was not related to presence of fluctuating Cooper pairs. It is also worth to notice that the effect was observed for samples thinner than ~15 nm and above this value films behaved like a bulk sample. In the present paper the 100 nm thick Ba(Fe1-xCox)2As2 layer is studied, which suggest that it should be treated rather as a 3D sample (especially as the anisotropy in this material is not large compared to, for example, cuprates).
My main complaint regarding the experimental part is that the authors present results obtained for just one sample, which was measured pristine and after ion bombardment. No dependence on the magnetic field or sample thickness is shown. Additionally, in the temperature range, where the authors claim the presence of superconducting fluctuations, I do not see any para-conductivity, which raises more doubts as to the interpretation given.
In conclusion, I think that without further support it is difficult to believe that the observed enhancement of Seebeck coefficient in Ba(Fe1-xCox)2As2 is due to fluctuating Cooper pairs. In addition, a main idea behind the studies has already been presented in Ref. [15]. Therefore, I recommend rejection of the manuscript.
At the end I have a few comments that perhaps might be useful for future submissions:
- The structure of the manuscript should be better organized.
- Abstract and introduction are way too long.
- Readability of Fig. 5 would benefit from changing scale from linear to logarithmic.

  • validity: -
  • significance: -
  • originality: -
  • clarity: -
  • formatting: -
  • grammar: -

Author:  Henni Ouerdane  on 2023-07-25  [id 3837]

(in reply to Report 1 on 2022-08-04)

First part of the comments by the Referee -- overall assessment and basic criticism

Our reply:

We thank the Reviewer for a very critical assessment of our work, which is useful , as we have sought to find a plausible interpretation of the experimental signatures we have found, especially the observed rise of the magnitude of the Seebeck coefficient starting near 50 K. In particular, we involved in these discussions a recognized expert on this material, Prof. Dr. Anna E. Boehmer, who is now also co-author of this paper. After intensive discussions, our interpretation is that the experimental signature in the Seebeck coefficient away from $T_{\rm c}$ could well be due to nematic fluctuations in the electron subsystem.

As a matter of fact, there is no universal model to simulate the thermoelectric properties of pnictide thin films across a wide range of temperatures. The approaches are essentially phenomenological. In our work we developed a thermodynamic model to discuss a particular experimental feature: the divergence of the power factor very close to $T_{\rm c}$, and in principle, the motif of fluctuations in the electron subsystem, at least in its phenomenology, translates very well from fluctuating Cooper pairs to nematic fluctuations. However, one has to say that for fluctuating Cooper pairs there have been established models for decades, whereas the mathematical description of nematic fluctuations is not yet at this established level. We have therefore chosen a compromise for the paper. In the theoretical description, we now make it clear that the crucial aspect comes from the fluctuation regime itself. In the experiment, we explain our data in terms of nematic fluctuations, which we substantiate with several references.

Next part of the comments by the Referee -- main complaint

Our reply:

We agree with the reviewer that consideration of fluctuating Cooper pairs alone here is insufficient. We have changed all text passages accordingly and interpret the data - in view of the literature - by nematic fluctuations.

In principle, we also agree with the reviewer that more experimental data would be necessary. However, this paper has a very strong theoretical ground. The experimental data should show that in principle the phenomenon of electronic fluctuations near phase transitions, shows up clearly in experimental transport data. A more detailed experimental study will follow in the future.

Third part of the comments by the Referee -- conclusion

Our reply:

The main idea behind Ref. [15] was to show quantitatively by means of numerical simulation of different models of electrically charged working fluids that if a system that can undergo a phase transition, its thermoelectric coupling strongly can be enhanced near the critical point. We illustrated this with the case of the 2D fluctuating Cooper pairs. The metric then introduced and used for comparison among different working fluids is the thermodynamic figure of merit $Z_{\rm th}T$, which is a combination of thermoelastic coefficients derived from the definition of the isentropic expansion factor $\gamma$.

In the present work, we significantly go beyond the scope of the previous work first by relating the thermodynamic figure of merit to the electronic thermoelectric figure of merit, i.e. the thermoelastic properties of the working fluid to its transport properties. To do so, we introduce an equivalent of the isentropic expansion factor based on transport coefficients, $\gamma_{\rm tr}$. The latter is a measure of the deviation from the thermostatic $\gamma$ when the system is out-of-equilibrium. In fact, thermoelectric conversion performance improves when $\gamma$ becomes large, which implies that the working fluid transport properties foster a high power factor if the working fluid's compressibility is larger. We see that $\gamma$ becomes large in the fluctuation regimes, and this is a nice illustration of the fluctuation-compressibility theorem in the context of thermoelectricity.

Further, contrary to Ref. [15], we present experimental data, although the development of a complete model that can addresses all aspects of the observed behaviors (Seebeck, electrical resistivity, power factor) in the clean and degraded samples is beyond the scope of the present work. What we had initially set out to do is to interpret and discuss the observed behavior close to $T_{\rm c}$ in light of our thermodynamic analysis, rather than provide a full model. It was clear that given the fact that we made a number of simplifying assumptions for clarity of the electronic working fluid models, we could not explain the trend starting at around 50 K down to close to $T_{\rm c}$. In the revised version, however, we provide a qualitative interpretation of the observed behavior in this range as a manifestation of nematic fluctuations.

We thus believe that in addition to the experimental data and their analysis, our thermodynamic study is not only non-trivial but also sheds light on the basic properties of the electron gas in thermoelectric materials. Further, we believe our work paves the way to the development of theoretical and experimental developments in the field of thermoelectricity as a number of questions remain to be answered.

Additional remarks by the Referee

our reply:

We have reshaped our manuscript to better report our work which has a large theoretical component.

True, the Introduction part is quite long, but we feel that it is necessary to properly lay out the rationale for this work, which is at the crossroads of various fields: thermodynamics, transport, solid-state physics, materials science, energy conversion. Same applies for the abstract that we tried to keep as concise yet as informative as possible.

Interestingly, the format of SciPost papers shows the content of the paper so any reader interested in particular aspects of the work can skip other parts that they find not essential for the understanding of the work. So, a long introduction provides more benefits than drawbacks here.

Close to $T_{\rm c}$ the rise is so sharp and the temperature range so small (to account for the limit of validity of the model) that a logarithmic scale does not provide a better view. In fact, this was our very first intent, but we decided that the linear scale provides a more satisfactory display of the curves.

Login to report or comment