Skip to main content
Open Research Europe logoLink to Open Research Europe
. 2023 Mar 7;3:43. [Version 1] doi: 10.12688/openreseurope.15367.1

The SquAd derivation: A Square Additive approach to the turbulent Prandtl number

Vincent Moreau 1,a
PMCID: PMC10873542  PMID: 38370030

Abstract

Liquid metals have been chosen as primary coolant of innovative nuclear systems under current development. They present a very high thermal conductivity and hence a very low molecular Prandtl number. This feature challenges the modeling of turbulent thermal flows applying the Reynolds analogy. This paper addresses this challenge. A new formula for the turbulent Prandtl number is derived in terms of local variables available from two-equations turbulence models. The derivation is a direct consequence of the expected square additivity of the molecular and turbulent parameters defining the effective viscosity and the effective conductivity. The formula does not degenerate and leads to a Kays like formulation if approximated. While constrained by the quality of the turbulent viscosity modeling, it has the potential to improve the numerical simulation of turbulent thermal flows.

Keywords: Prandtl, Kays, CFD, turbulence modelling

Introduction

Low Prandtl number liquid metals serve as primary coolant for MYRRHA 1 and ALFRED 2 , two Gen-IV reactors under development 1 . The low Prandtl number induces discrepancies in the modeling of the turbulent heat transfer when directly addressed according to the Reynolds analogy with a constant turbulent Prandtl number. The thermal boundary layer is considerably larger than the velocity one, up to the point that, while we can clearly define a bulk velocity, the temperature profile may not exhibit any almost constant bulk temperature plateau. The issue was investigated by several authors who reviewed the existing correlations and proposed their own one 24 . The correlations make use of adimensional numbers such as Reynolds and Peclet. Their functional form is mostly empirical with coefficients determined by best fit. The Reynolds number and the Peclet numbers are global parameters, only useful and well-defined for simple geometries. Their use in CFD 3 simulations of complex geometries is questionable. Among all correlations reviewed, Kays’ correlation is the only one making exclusive use of local parameters. Variants of this correlation have been used with significant success. The variants share the same structure but feature different values for one numerical constant. In our previously published four page brief report 5 , we showed that the correlations can be simply derived on a basic assumption with regards to the non-linear combination of stochastic effects and the variants then stem from different approximations of a mother formula. The objective of this former brief report was only to establish and keep trace of the probable relationship between the stochastic concept and the empirical correlation. It is in no way a bulletproof demonstration and there is no specific treatment of the viscosity. In particular, the definition of the asymptotic Prandtl number and its use are not completely consistent. Besides, from the purely numerical point of view, a defect of the mother formula, which is transferred to the variants, is that the turbulent Prandtl number becomes infinite at vanishing turbulence.

These considerations motivate to proceed further with the analysis which is the object of this current brief report. The key point resides in refining the concept that lies behind the formerly loosely determined asymptotic Prandtl number. What was really needed to be done was apparently to clearly separate what is from molecular origin and what is from turbulent origin. This was found to be more prolifically reformulated in terms of differentiating the properties of the fluid from the properties of the flow, being this later one either laminar or turbulent.

In order to proceed consistently with the flow versus fluid properties separation it becomes necessary to apply the principle of square additivity not only to the thermal conductivity but also to the viscosity. This allows us to have a cleaned definition of what is now called the ”flow Prandtl number” which is conceived to be a universal constant of the flow, independent of the fluid. As previously, the principle of square additivity completely determines the turbulent Prandtl number. The new formula, while similar, is more articulated than the previous one presented. Its first order approximation still has the functional shape of Kay’s correlation. It however gives better insight on its probable range of validity. Besides, as a good news for numerical implementation, the newly derived formula has the merit of not degenerating anymore at vanishing turbulence, something which was not a feature of the previous formula.

Turbulent prandtl number derivation

In the the framework of thermal fluid dynamics of turbulent flows, the focus is concentrated on the determination and modelling of the effective viscosity and effective heat diffusion of the fluid. The viscosity and the heat diffusion both consist in the sum of two parts, one molecular and the other turbulent. This is a convenient and simple representation based on the fact that for most turbulent flows, the turbulent viscosity and conductivity are much higher than their molecular counterparts. The turbulent quantities were also not expected, at least historically, to be known at a high level of precision.

With regard to the heat diffusion, both the molecular and the turbulent fluxes are oriented in the direction of the local temperature gradient and proportional to it. The intensity is also determined by conductivity coefficients. The effective conductivity coefficient k e can be expressed as

ke=k+kt(1)

denoting, respectively with k and k t the molecular conductivity and the increment due to turbulence.

Indicating with ρ the fluid density and C p its specific heat, this expression can be rewritten in terms of thermal diffusivity:

αe=α+αt(2)

in which α e = k e /( ρC p ) stands for the effective diffusivity, α = k/( ρC p ) is the molecular part and α t = k t /( ρC p ) is the turbulent part, ρ is the density and C p the specific heat.

Similarly, as far as the viscosity in concerned, the effective kinematic viscosity ν e is the sum of a laminar contribution ν and a turbulent contribution ν t :

νe=ν+νt(3)

The argument previously developed in 5 is the following. Conduction has a stochastic origin. The scale at which molecular and turbulent conduction operate are different and their mechanisms are unrelated. Molecular conduction is a molecular process and is a property of the fluid while turbulent conduction has a convective origin and is a property of the flow. The combination of their effects is better represented as a convolution rather than as a direct sum. Translated in formula, we expect to have:

αe=α2+α02,(4)

where the turbulent diffusivity α 0 should be an intrinsic property of the flow rather than the fluid.

The same argument is extended to the viscous process and states that the effective viscosity comes from two independent processes whose intensity should be square additive:

νe=ν2+ν02.(5)

By simple substitution, we have just defined two quantities:

ν0=νt1+2ννt(6)

and

α0=αt1+2ααt.(7)

We can also redefine ν t and α t in terms of these quantities:

νt=ν2+ν02ν(8)
αt=α2+α02α(9)

We now introduce the Prandlt number Pr=να , the turbulent Prandtl number Prt=νtαt and the flow Prandtl number as Pr0=ν0α0 . This later number has been conceived in order to be independent of the fluid and only dependent on the flow. Noting that αα0=Pr0νPrν0 , by combining the former equations, we obtain:

Prt=Pr02Pr1+(Prν0Pr0ν)2+11+(ν0ν)2+1.(10)

This specific form is chosen to show that it cannot degenerate to zero or infinity. In effect, because ν 0 goes to zero if ν t does, then Pr t tends to Pr02Pr while for large turbulence, Pr t tends to Pr 0. In turn, if Pr 0, built for this purpose, does not degenerate, then neither does Pr t .

The former expression can be rewritten in terms of ν/ ν 0:

Prt=Pr01+(Pr0νPrν0)2+Pr0νPrν01+(νν0)2+νν0.(11)

This formula is quite complicated and not easy to interpret at first glance, but considering ν/ ν 0 as a small parameter, a brutal simplification of 11 at first order in this parameter gives:

PrtPr0[1+(Pr0Pr1)νν0](12)

This formulation is practical only if ν 0 is readily available. This would be the case if we had a transport equation for ν 0 or a related variable similarly to what is done with the usual 2-equations turbulence models. Exploring the potential of this possibility is beyond the scope of the current argument and we need to express Pr t in terms of known parameters.

Thus, expressing Pr t in terms of ν t instead of ν 0, the formula becomes:

Prt=Pr02Pr1+(PrPr0)2νtν(νtν+2)+12+νtν,(13)

this form being useful for interpretation at vanishing turbulent viscosity.

Dividing this last formula by Pr 0, we have a relation between the three adimentional groups PrtPr0 , PrPr0 and νtν in the form:

PrtPr0=f(PrPr0,νtν).(14)

In view of a development in ν/ ν t , the formula becomes:

Prt=Pr01+2ννt+(Pr0νPrνt)2)+Pr0νPrνt1+2ννt(15)

Here again, a brutal first order development in terms of ν/ ν t , valid only when both ν t / ν and Prν t / Pr 0 ν are large, gives:

PrtPr0[1+(Pr0Pr1)ννt],(16)

meaning that changing from ν 0 to ν t brings only second order terms approximation. In particular, this last expression has the same form as the Kays correlation 2 .

Discussion

Under the hypothesis that the flow Prandtl number Pr 0 is constant, it coincides with the value used for highly turbulent flow and near unit Pr. That is, Pr 0 = 0.85. We get for Pr ≃ 0.025 typical of heavy liquid metal:

Prt0.85+0.70Prνtν(17)

For the second coefficient (here 0.7), Kays indicated two values, 0.7 and 2, discussing without reaching a conclusion in favor of one or the other value. However, a less brutal approximation could lead to a different coefficient. For example, in 6, 1.46 is obtained from direct analytical integration and best fit of heat transfer in a tube. Moreover, our newly derived formula is not much different than the one derived previously 5 and a more precise approximation, while more complex to derive, would most probably also lead to an increased second coefficient about 1.45.

There are indications 2 that Pr t and therefore also Pr 0 is about 0.85 for very large ν t / ν. We do not know the behavior of Pr 0 for lower values. Nevertheless, all this construction makes sense only under the hypothesis of a basically constant Pr 0. A particular case is when Pr = Pr 0. Then Pr t = Pr 0 too. For medium and high Pr, we observe that Pr t in formula ( 11) does not significantly departs from Pr 0, except for strongly vanishing ν t / ν.

With a little algebra, we found that Pr t is a decreasing function of ν t for PrPr 0 and increasing function of ν t for PrPr 0, with values spanning the interval [ Pr 0; Pr02 / Pr] and [ Pr02 / Pr; Pr 0]. The approximation 16, while with the same monotonicity, fails to meet the correct bound for vanishing viscosity, for which it degenerates. Interestingly, we can see that the second coefficient depends critically on the Prandtl number, to the point that it vanishes for Pr t = Pr 0 and changes sign afterwards, like for the complete expression. This is a clear indication that the Kays correlation should be used as such only for low Prandtl number (say < 0.1) fluids.

The derived formula is ought to be used within a turbulence model. It would make sense only if the turbulence model correctly predicts the turbulent viscosity. This has to be true not only in the viscous boundary layer but also and principally in the bulk, in which the thermal boundary layer could still be in development. The problem is that the turbulence models mainly focus on the correct near-wall boundary layer turbulent viscosity profile, as it is the place where almost all the pressure drop is built, with the main aim to capture the correct wall shear stress. The turbulent viscosity profile in the bulk is normally of no practical importance, except for thermal flows of low Prandtl fluids. Looking at the profile of ν t compared with direct numerical simulation data even in the simpliest 2D channel flow, see figure 5 in 7, we can see that the turbulence models fail to give a correct profile for a wall Y+ above 30. In particular, ν t is underpredicted below Y+ up to 100-150 and over predicted afterward. The difference between the simply additive and the square additive approaches is mainly concentrated and felt around the values where both contributions are of similar intensity: ν/ Prν t / Pr 0 or equivalently ν t / νPr 0/ Pr. For a low Prandtl fluids with Pr = 0.025 we have Pr 0/ Pr = 34, so all the region where ν t / ν lies between say 10 and 100 is concerned, that is precisely for Y+ above 30 for the case analysed in 7. The balance between the thermal effects of both Y+ region is shifted towards an artificially increased diffusion. To counteract this effect, the turbulent Prandtl number can be increased artificially for a better fit. This could be an explanation for the use of an augmented second coefficient in the Kays correlation as indicated previously.

The main effect of applying the square additivity to the effective viscosity is that it removes the degeneration of the Pr t formula for vanishing viscosity that was still present in 5. It is not clear nor sure that it is of practical importance anywhere else. It seems that, within the level of approximation given by the 2-equations turbulence models, ν t and ν 0 can be used indifferently in the formula. In other words, the square additivity could be used solely for the energy equation.

Conclusions

We consider that turbulent viscosity and molecular viscosity are originated from two independent stochastic processes whose intensity is square additive. We make the same consideration for the turbulent conductivity and the thermal conductivity. We have defined a flow Prandtl number which is expected to be a universal property of the flow and to be in fact a constant under the square additive approach. The Prandtl number is then determined by a formula reproducing the first variant of Kays’ correlation when approximated at the first order. The derivation sheds light on the Kays correlation and indicates that the second coefficient depends critically on the Prandtl number to the point that it vanishes when Pr = 0.85. Under the condition that the classical 2-equations turbulence models become able to capture correctly the turbulent viscosity profile, we expect that directly implementing the square additivity of the components of the effective conductivity could give improved thermal results independently of the Prandtl number and particularly for the low Prandtl liquid Lead and Lead alloys.

The Reynolds analogy and its extension to thermal flows could have a much wider domain of validity than expected by combining it with the square additivity of the coefficients.

Acknowledgments

A previous (non-peer reviewed) version of this article is available on preprints.org ( https://www.preprints.org/manuscript/202212.0160/v1) and the publications.crs4.it repository ( http://publications.crs4.it/pubdocs/2022/Mor22/).

Funding Statement

This research was financially supported by the European Union’s Euratom programme under the grant agreement No 945077 (Partitioning And Transmuter Research Initiative in a Collaborative Innovation Action [PATRICIA]).

[version 1; peer review: 1 approved, 1 not approved]

Footnotes

1 Multi-purpose Hybrid Research Reactor for High-tech Applications

2 Advanced Lead-cooled Fast Reactor European Demonstrator

3 Computational Fluid Dynamics

Data availability

No data are associated with this article.

References

  • 1. Tarantino M, Roelofs F, Shams A, et al. : SESAME project: advancements in liquid metal thermal hydraulics experiments and simulations. EPJ Nuclear Sci Technol. 2020;6(8):18. 10.1051/epjn/2019046 [DOI] [Google Scholar]
  • 2. Kays WM: Turbulence Prandtl Number - Where are we? J Heat Transfer. 1994;116(2):284–285. 10.1115/1.2911398 [DOI] [Google Scholar]
  • 3. Cheng X, Tak NI: Investigation on turbulent heat transfer to lead- bismuth eutectic flows in circular tubes for nuclear applications. Nucl Eng Des. 2006;236(4): 385-393. 10.1016/j.nucengdes.2005.09.006 [DOI] [Google Scholar]
  • 4. Bartosiewiccz Y, Duponcheel M, Mancono M, et al. : Turbulence Modeling at Low Prandtl Number. Chapter in Fluid Mechanics and Chemistry for Safety Issuest in HLM Nuclear Reactors. Lecture Series 2014–02 by the von Karman Institute for Fluid Dynamics. [Google Scholar]
  • 5. Moreau V: Turbulent Prandtl Number, reformulation of Kay’s correlation. Academia Letters. 2021;2366. 10.20935/AL2366 [DOI] [Google Scholar]
  • 6. Taler D: Heat transfer in turbulent tube flow of liquid metals. Procedia Eng. IX International Conference on Computational Heat and Mass Transfer, ICCHMT2016.2016;157:148–157. 10.1016/j.proeng.2016.08.350 [DOI] [Google Scholar]
  • 7. Oder J, Koloszar L, Fiore M, et al. : Report on the comparison of the explicit AHFM with available test cases and on the implementation of the implicit AHFM in OpenFOAM. Deliverable D11.5, PATRICIA European Project Grant Agreement Number 945077,2022. [Google Scholar]
Open Res Eur. 2023 Jun 8. doi: 10.21956/openreseurope.16612.r32061

Reviewer response for version 1

Davide Modesti 1

The manuscript introduces the concept of "flow Prandtl number", namely a universal constant that does not depend on the fluid. This concept is introduced theoretically but then it is not verified using available experimental or numerical data, which is a major shortcoming of this work.

I believe that the manuscript has the following shortcomings:

  1. I don't understand why we need to introduce a flow Prandtl number. Isn't it enough to say that the turbulent Prandtl number is independent from the molecular Pr?

  2. My main concern is that conclusions are not supported by results. After the derivation of the expression for the flow Prandtl number, I would have expected a validation of the theory using available numerical data. There are many data on this topic. For instance, we have recently generated a comprehensive dataset of turbulent plane channel flow spanning Prandtl numbers between 0.0025-4, which also covers the low Prandtl number regime of this manuscript. The manuscript is available on JFM (Pirozzoli and Modesti, 2023 1 ) and the data are also available online at http://newton.dma.uniroma1.it/. 

  3. The other major concern that I have is on the effective diffusivity in equation (4). The canonical definition of thermal diffusivity (equation 2) is not arbitrary, but it stems from the mean temperature equation upon introducing the eddy diffusivity hypothesis. How does equation (4) comply with the mean temperature balance?

  4. The new formulation should be validated against reference numerical data and other state of the art formulas for the turbulent Prandtl number, such as the one by Cebeci (1973 2 ), or the variant proposed by Na and Habib (1973 3 ), or the formula by Kays et al. (1980 4 ).

  5. In the text of the manuscript I see reference to figures, but I cannot visualize any figure.

Is the study design appropriate and does the work have academic merit?

Partly

Is the work clearly and accurately presented and does it cite the current literature?

Partly

If applicable, is the statistical analysis and its interpretation appropriate?

Not applicable

Are all the source data underlying the results available to ensure full reproducibility?

No source data required

Are the conclusions drawn adequately supported by the results?

No

Are sufficient details of methods and analysis provided to allow replication by others?

Yes

Reviewer Expertise:

wall turbulence; turbulent flows; compressible flows; forced thermal convection; heat transfer

I confirm that I have read this submission and believe that I have an appropriate level of expertise to state that I do not consider it to be of an acceptable scientific standard, for reasons outlined above.

References

  • 1. : Direct numerical simulation of one-sided forced thermal convection in plane channels. Journal of Fluid Mechanics .2023;957: 10.1017/jfm.2023.104 10.1017/jfm.2023.104 [DOI] [Google Scholar]
  • 2. : A Model for Eddy Conductivity and Turbulent Prandtl Number. Journal of Heat Transfer .1973;95(2) : 10.1115/1.3450031 227-234 10.1115/1.3450031 [DOI] [Google Scholar]
  • 3. : Heat transfer in turbulent pipe flow based on a new mixing length model. Applied Scientific Research .1973;28(1) : 10.1007/BF00413075 302-314 10.1007/BF00413075 [DOI] [Google Scholar]
  • 4. : Convective Heat and Mass Transfer. McGraw-Hill, New York .1980;
Open Res Eur. 2023 Jun 21.
Vincent MOREAU 1

The reviewer text is indicated in italics.

The manuscript introduces the concept of "flow Prandtl number", namely a universal constant that does not depend on the fluid.

That is correct.

This concept is introduced theoretically but then it is not verified using available experimental or numerical data, which is a major shortcoming of this work.

The concept is indeed first introduced theoretically together with the necessary notation. The notation has been clarified giving proper names to α 0 and ν 0. The subscript 0 is not very satisfying but is raised necessary for consistency with the definition of Pr 0 for which it makes sense. For a comparison with published data it is necessary to look at the consequences of the hypothesis on Pr t expressed in terms of the eddy viscosity, eq. 13 which is the most practical form for numerical implementation but not for further understanding of its meaning. Before proceeding with the validation, eq. 13 is rewritten as eq. 15 to make appear clearly the near constant aspect perturbed by the “natural” small parameter nu/nu t.

The concept is verified from experimental data at the very beginning of the Discussion section. In effect, eq.17 (first Kays correlation in ref.2) is retrieved by substantiation of eq.16 for Pr=0.025. As in turn eq. 16 is a first order approximation of eq. 15, put in this form for this purpose. From what I have extensively seen and read, for liquid metals, Kays’ correlation is the ultimate comparison to overcome for more elaborated turbulence thermal models, such as 3 or 4 eq. models.

So, I see that my formula is substantially identical, up to a second order term to Kays’ correlation and I sincerely think that it is more than enough for a validation when dealing with liquid metals which is the main focus of this brief report. Besides,  my derivation gives better insight on the range of validity of the Kays correlation. This can help users avoid using it for about unit Pr t. As easily seen from formula 13 or 15, when Pr=Pr_0 then Pr_t becomes constant which is the choice made by the main CFD code providers with their extensive suite of validation. I have seen the constant taken either at 0.85 or 0.9. It means that there is already there a non negligible uncertainty. This is also why I do not think that going above the first order approximation for validation makes sense.

I believe that the manuscript has the following shortcomings:

I don't understand why we need to introduce a flow Prandtl number. Isn't it enough to say that the turbulent Prandtl number is independent from the molecular Pr?

This is all the point of this brief report. While the hypothesis you state is usually sufficient for engineering application it is no more when dealing with liquid metals. In this case, the region where nu and nu_t are of similar order is much wider that usual and simple summation leads to an overestimation of the effective heat transfer. Would it have been an underestimation, things would have been easier because a constant Pr_t would have provided a slightly conservative solution. Modifying the constant value can be done a posteriori but is not satisfying for prediction capabilities. Besides, we commercial CFD users have the pretention not only to give one or two global parameters but also to give better global fields description. So, it is necessary at least, and still waiting for better, to use Kays’ correlation. Problem is that there are two such correlations and authors have also interpolated between them. We are still stuck with the difficulty to reach predictability of the results.

By giving a theoretical foundation to the Kays correlation which address the issue of the dependence of Pr_t on Pr, I believe this brief node can give a useful contribution.

 

My main concern is that conclusions are not supported by results. After the derivation of the expression for the flow Prandtl number, I would have expected a validation of the theory using available numerical data. There are many data on this topic. For instance, we have recently generated a comprehensive dataset of turbulent plane channel flow spanning Prandtl numbers between 0.0025-4, which also covers the low Prandtl number regime of this manuscript. The manuscript is available on JFM (Pirozzoli and Modesti, 20231) and the data are also available online at http://newton.dma.uniroma1.it/.

This point has been abundantly answered before. The validation comes directly from the extensively validated Kays’ correlation from the one hand and from the constant value when Pr=Pr_0 fro the other hand.

 

The other major concern that I have is on the effective diffusivity in equation (4). The canonical definition of thermal diffusivity (equation 2) is not arbitrary, but it stems from the mean temperature equation upon introducing the eddy diffusivity hypothesis. How does equation (4) comply with the mean temperature balance?

I have named α 0 and ν 0 to clarify the procedure. The description with (α 0, ν 0) and (α t, ν t) are completely equivalent and can be retrieved one another. The first couple comes from physical/stochastic consideration and the second couple from the averaging procedure of the NS equations. They describe the same thing with a different couple of variables. It is only a reversible change of variable from eq. 4 to eq. 9. You can also see (alpha_0, nu_0) as intermediate variables aimed at a simple definition of Pr_0, somewhat like imaginary numbers have been used for so many years as a trick to factorize 3dt and 4 th order polynomials.

 

The new formulation should be validated against reference numerical data and other state of the art formulas for the turbulent Prandtl number, such as the one by Cebeci (19732), or the variant proposed by Na and Habib (19733), or the formula by Kays et al. (19804).

The new validation is indeed validated against the formula by Kays but taking his much more recent, even if already quite old, comprehensive paper of 1994, see. Ref. 2, which is precisely the core of the validation.

 

In the text of the manuscript I see reference to figures, but I cannot visualize any figure.

The reference to a figure is given in the penultimate paragraph of the Discussion section. It is written “see figure 5 in 7” where 7 is the number for the reference and in effect the figure is not in the brief report. There is a very large number of figures in the referenced document and I thought it would ease the reader who would like to look at the source and not only rely on the description following the reference.

This reference is particularly important because before its publication I could not find, probably my bad, any ν t profile usable for interpretation and/or reference. I indeed postponed the submission of this brief report until the reference 7 is at least theoretically publicly available.

My first brief report submission was much more concise, but at the editorial level I was asked to develop some arguments and then new requests arise about the new development in something that looked like and endless loop. Most of the Discussion section arises from this interaction. I had to put a stop to it, because I want this brief report to remain a brief report. It is also why I do not want to add any graph or picture. That would restart the endless loop without adding substance. If one point must be kept after the validation part, it is that it would be very easy to discard my finding, independently of its validity, by testing it against some of the popular 2-eq models, just because my formula relies on the ν t profile and this ν t profile is very badly reproduced, as shown in ref.7. Unfortunately, even if valid, the new formula do not lead directly to much simulation result improvement. However, it suggests to work on the existing 2-equations turbulence models so that they can procure valid ν t profiles and not only the correct pressure loss.

As a final word, please also look at the “Amendment from Version 1”. Looking at the brief report after a few month, let’s say with refreshed eyes, I could see and correct many imprecisions and/or incorrect wording. I hope the version 2 of the document is now more clear and easier to understand.

Open Res Eur. 2023 May 30. doi: 10.21956/openreseurope.16612.r32057

Reviewer response for version 1

Giacomo Barbi 1, Lucia Sirotti 2

The present article deals with the analysis of an explicit formula for the turbulent Prandtl number (Pr t) evaluation. The use of a square additivity approach leads to a new formulation that is coherent with other well-known expressions from turbulence modeling literature. The author extends a previous result, improving upon it by avoiding the degeneration of Pr t to zero or infinity. Moreover, the author's formula can be related to the Kays correlation by considering a first-order development of the full new expression.

Overall, this work is interesting for thermal turbulence modeling, and the manuscript is well-organized. However, there are a few issues that need clarification:

  • Since the meaning of Pr 0 can be seen as the asymptotic value of Pr t, it is suggested to underline the meaning of α 0 and ν 0. It should be noted that these properties are introduced with the same description as α t and ν t, representing properties of the flow rather than the fluid.

  • In formula 15 a bracket is missing.

  • It is recommended, if possible, to include a plot depicting the behavior of the new formulation of Pr t. The plot should consider a comparison with the Kays formula and highlight the different situations described in the first paragraphs of the Discussion section.

Is the study design appropriate and does the work have academic merit?

Yes

Is the work clearly and accurately presented and does it cite the current literature?

Yes

If applicable, is the statistical analysis and its interpretation appropriate?

Not applicable

Are all the source data underlying the results available to ensure full reproducibility?

No source data required

Are the conclusions drawn adequately supported by the results?

Partly

Are sufficient details of methods and analysis provided to allow replication by others?

Yes

Reviewer Expertise:

Turbulence modeling of low Prandtl number fluid. Finite element implementation of thermal turbulence models for liquid metal.

We confirm that we have read this submission and believe that we have an appropriate level of expertise to confirm that it is of an acceptable scientific standard.

Open Res Eur. 2023 Jun 6.
Vincent MOREAU 1

Thank you very much for your fast and accurate review. About the 3 issues raised:

  1. You are totally right. I struggled a lot to make things clear without so much success. What I intend to do for improvement is to names the quantities more accurately:
    • remove "turbulent" after eq. 4
    • add "the flow viscosity nu_0" before eq. 6
    • add "the flow diffusivity alpha_0" before eq. 7
  2. In formula 15, the last bracked must indeed be removed

  3. I would definitively prefer not to add a plot. This is a brief report and is already too much extended for my taste, as I had to compose with the editorial team. What you recommend, while surely of interest, would bring this brief report outside of its domain of definition.

I will wait for another review before making an update. Thank you again.

Associated Data

    This section collects any data citations, data availability statements, or supplementary materials included in this article.

    Data Availability Statement

    No data are associated with this article.


    Articles from Open Research Europe are provided here courtesy of European Commission, Directorate General for Research and Innovation

    RESOURCES