The Stanton number, St, is a dimensionless number that measures the ratio of heat transferred into a fluid to the thermal capacity of fluid. The Stanton number is named after Thomas Stanton (engineer) (1865–1931).[1][2]: 476  It is used to characterize heat transfer in forced convection flows.

Formula edit

 

where

It can also be represented in terms of the fluid's Nusselt, Reynolds, and Prandtl numbers:

 

where

The Stanton number arises in the consideration of the geometric similarity of the momentum boundary layer and the thermal boundary layer, where it can be used to express a relationship between the shear force at the wall (due to viscous drag) and the total heat transfer at the wall (due to thermal diffusivity).

Mass transfer edit

Using the heat-mass transfer analogy, a mass transfer St equivalent can be found using the Sherwood number and Schmidt number in place of the Nusselt number and Prandtl number, respectively.

 [4]

 [4]

where

  •   is the mass Stanton number;
  •   is the Sherwood number based on length;
  •   is the Reynolds number based on length;
  •   is the Schmidt number;
  •   is defined based on a concentration difference (kg s−1 m−2);
  •   is the velocity of the fluid

Boundary layer flow edit

The Stanton number is a useful measure of the rate of change of the thermal energy deficit (or excess) in the boundary layer due to heat transfer from a planar surface. If the enthalpy thickness is defined as:[5]

 

Then the Stanton number is equivalent to

 

for boundary layer flow over a flat plate with a constant surface temperature and properties.[6]

Correlations using Reynolds-Colburn analogy edit

Using the Reynolds-Colburn analogy for turbulent flow with a thermal log and viscous sub layer model, the following correlation for turbulent heat transfer for is applicable[7]

 

where

 

See also edit

Strouhal number, an unrelated number that is also often denoted as  .

References edit

  1. ^ Hall, Carl W. (2018). Laws and Models: Science, Engineering, and Technology. CRC Press. pp. 424–. ISBN 978-1-4200-5054-7.
  2. ^ Ackroyd, J. A. D. (2016). "The Victoria University of Manchester's contributions to the development of aeronautics" (PDF). The Aeronautical Journal. 111 (1122): 473–493. doi:10.1017/S0001924000004735. ISSN 0001-9240. S2CID 113438383. Archived from the original (PDF) on 2010-12-02.
  3. ^ Bird, R. Byron; Stewart, Warren E.; Lightfoot, Edwin N. (2006). Transport Phenomena. John Wiley & Sons. p. 428. ISBN 978-0-470-11539-8.
  4. ^ a b Fundamentals of heat and mass transfer. Bergman, T. L., Incropera, Frank P. (7th ed.). Hoboken, NJ: Wiley. 2011. ISBN 978-0-470-50197-9. OCLC 713621645.{{cite book}}: CS1 maint: others (link)
  5. ^ Crawford, Michael E. (September 2010). "Reynolds number". TEXSTAN. Institut für Thermodynamik der Luft- und Raumfahrt - Universität Stuttgart. Retrieved 2019-08-26.
  6. ^ Kays, William; Crawford, Michael; Weigand, Bernhard (2005). Convective Heat & Mass Transfer. McGraw-Hill. ISBN 978-0-07-299073-7.
  7. ^ Lienhard, John H. (2011). A Heat Transfer Textbook. Courier Corporation. p. 313. ISBN 978-0-486-47931-6.