Skip to main content

Mechanism of Bubble Formation

  • Chapter
  • First Online:
Vesiculation and Crystallization of Magma

Part of the book series: Advances in Volcanology ((VOLCAN))

  • 641 Accesses

Abstract

In the previous chapter, we learned that, for bubble generation, a supersaturated state is required. So, how do bubbles actually form in a supersaturated liquid? How do these bubbles grow to bubbles that we can see? To answer these questions and understand bubble formation, it is absolutely necessary for us to consider the meaning of the size and number of bubbles present in highly vesiculated volcanic rocks such as pumice and scoria, described in Chap. 1. These concepts are necessary for us to understand the dominant factors controlling eruption styles and intensity from the bubble texture in highly vesiculated volcanic rocks. In this chapter, we will understand the mechanisms of such bubble formation and behavior of bubbles in liquids based on both theories and experiments.

This is a preview of subscription content, log in via an institution to check access.

Access this chapter

Chapter
USD 29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD 129.00
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 169.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info
Hardcover Book
USD 169.99
Price excludes VAT (USA)
  • Durable hardcover edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

Notes

  1. 1.

    The word stable used in this sentence actually means metastable. Water can be stable as liquid water even if it is cooled down to 0 \(^{\circ }\!\)C, i.e., the freezing point. This state is a metastable state and such water is sometimes called supercooled water. Let us consider a metastable state from a point of view of energy under gravity as an analogy. When you place a cuboid on a horizontal plane, you can do it in two ways: where the longest sides are perpendicular to the horizontal plane and where the longest sides are parallel to the horizontal plane. Although the latter has lower potential energy and is stable, the former is stable unless pressed by an external force that is larger than a certain degree of force.

  2. 2.

    This bubble is a very small one comprising 100–1000 water molecules.

  3. 3.

    The number of possible states indicates the number of possible combinations, under a given total energy of a designated system, of the arrangement of different energy states that microscopic parts constituting the system contain. The maximum value of it corresponds to the arrangement that has the most ways of combination, i.e., the case that has the maximum arrangement of microscopic energy states is an equilibrium state, and entropy at the state reaches a maximum, given by Eq. (3.1).

  4. 4.

    Conventionally, the probability of nucleation has been evaluated by translating internal energy into work (reversible work) under constant entropy and using it based on Landau and Lifshitz (1958). This book uses free energy to determine direct connection to calculations that are handled later.

  5. 5.

    The mark \(*\) is used as a mark representing saturation states.

  6. 6.

    A critical nucleus is a bubble nucleus or a crystal nucleus having a critical radius, which is called the critical nucleus radius or simply the critical radius, i.e., the radius of a critical nucleus.

  7. 7.

    Here it is not definitely stated whether the number of bubbles at the critical size is a value in the equilibrium size distribution \(N_\mathrm{E}^{*}\) or a value in the stationary size distribution \(N_\mathrm{S}^{*}\) discussed later. They are in proportion to each other, as seen later.

  8. 8.

    Note that the migration velocity used here is different from the growth rate. The migration velocity used here includes the effect of fluctuation. \(F^{*}\) can be represented by the equilibrium size distribution \(F_\mathrm{E}^{*}\). Even if the stationary distribution \(F_\mathrm{S}\) is assumed, the flux J of clusters with fluctuation in the size space cannot be simply expressed by the product of size distribution \(F_\mathrm{S}\) and the growth rate A (actual rate at which clusters at the critical size grow, i.e., zero), \(A \cdot F_\mathrm{S}\). J should have the fluctuation term (B), i.e., \(J=A \cdot F - B \cdot \partial F / \partial R\) (Eq. 3.34). It is rather preferable that \(F^{*}\) is symbolically represented by a value of the equilibrium size distribution \(F_\mathrm{E}^{*}\) at the critical size.

  9. 9.

    This is an expedient approach to hold change in size associated with coalescence and fragmentation of bubbles and conservation of mass and to simplify an expression of the addition and subtraction of the size. It will be converted to the radius of bubble later. Assuming that the molecular volume of gas is \(v_\mathrm{G}, l v_\mathrm{G} = 4 \pi R^{3}/3\).

  10. 10.

    Therefore, this equation can be universally applied to other aspects such as change in populations of human beings or insects by replacing the size in Eq. (3.27) with a human being’s income or the size of the insects. However, it is possible only when the transition probability can be completely described as a function of size.

  11. 11.

    See Sect. 11.4 in the Appendix (Chap. 11). As the ambiguity concerning \(F_\mathrm{E}^{0}\) remains to the end, it should be accurately described from the microscopic point of view. Regarding homogeneous reaction of solution and gas, a partition function of the transition state is statistically defined and the transition rate is calculated by the theory of the transition state.

  12. 12.

    This is the approach of Lifshitz and Pitaevskii (1981) and is convenient because the result is determined only by well-known macroscopic physical properties. Another method is to calculate a reaction rate in accordance with the theory of the transition state in terms of statistical mechanics.

  13. 13.

    Toramaru (1995) introduces \(y=C_\mathrm{eq}/C\) to obtain \(R_{0} = 2 \gamma /P \cdot P/P_\mathrm{G}^{*}= 2 \gamma y^{2} /P \) and defines \(v_\mathrm{G}\) by \(v_\mathrm{G}= k_\mathrm{B} T/P\). In this case,

    $$\begin{aligned} B(R) =\frac{v_\mathrm{G}^{2} D Cy^{2} }{8 \pi R_\mathrm{C} R^{2}} \end{aligned}$$
    (3.74)

    is obtained, which contains \(y^2\) in its numerator. Note that \(v_\mathrm{G}\) is the molecular volume of gas under the liquid pressure P.

  14. 14.

    If the coefficient of the equilibrium size distribution \(F_\mathrm{E}^{0}\) is left as it is,

    $$\begin{aligned} J_\mathrm{S} = Z \frac{v_\mathrm{G}^{2} D C F_\mathrm{E}^{0}}{8 \pi R_\mathrm{C}^{3}} \exp \left( - \displaystyle \frac{4 \pi \gamma R^{2}_\mathrm{C} }{3 k_\mathrm{B} T } \right) \end{aligned}$$
    (3.81)

    is.

  15. 15.
    $$\begin{aligned} J_\mathrm{S} = \frac{D C^{2} ( 1-y^{2} ) }{8 \pi } \sqrt{\displaystyle \frac{k_\mathrm{B} T}{\gamma }} \exp \left( - \displaystyle \frac{4 \pi \gamma R^{2}_\mathrm{C} }{3 k_\mathrm{B} T } \right) = Z \frac{v_\mathrm{G} D C^{2} y^{2}}{8 \pi R_\mathrm{C}} \exp \left( - \displaystyle \frac{4 \pi \gamma R^{2}_\mathrm{C} }{3 k_\mathrm{B} T } \right) \end{aligned}$$
    (3.83)

    where \(v_\mathrm{G}=k_\mathrm{B} T / P\text {D}\) .

  16. 16.

    Force balance in a direction perpendicular to the interface between the crystal and liquid is compensated by the strength of the crystal.

  17. 17.

    \(c_{n}\) is used rather than c because there can be multiple constants.

  18. 18.

    This equation is a typical second-order ordinary differential equation

    $$\begin{aligned} \frac{d^{2} y(x)}{dx^{2}} + p(x) \frac{dy(x)}{dx} + q(x) y(x) =0 \end{aligned}$$
    (3.108)

    in which the coefficient is not a constant. Many equations appearing in mathematical physics, such as the Legendre equation and Bessel’s equation, result in this form. Because the case involving constant coefficients is familiar, the case in which coefficients are not constants seems to be easily solved. However, the normal method of solving it is very complicated.

  19. 19.

    This solution is expressed by the result in which solutions to an infinite number of eigenvalues are superimposed.

References

  • Becker R, Döring W (1935) Ann d Physik [5] 24:719

    Google Scholar 

  • Blander M, Katz JL (1975) Bubble nucleation in liquids. AIChE J 21(5):833–848

    Article  Google Scholar 

  • Cluzel N, Laporte D, Provost A, Kannewischer I (2008) Kinetics of heterogeneous bubble nucleation in rhyolitic melts: implications for the number density of bubbles in volcanic conduits and for pumice textures. Contrib Mineral Petrol 156:745–763

    Article  Google Scholar 

  • Friedlander SK (1977) Smoke, dust and haze: fundamentals of aerosol behaviour. Wiley

    Google Scholar 

  • Haken H (1978) Synergetics: an introduction. Non-equilibrium phase transition and self-self organization in physics, chemistry and biology, 2nd ed. Springer

    Google Scholar 

  • Hirth JP, Pound CM, Pierre GR St (1970) Bubble nucleation. Metal Trans 1:939–945

    Google Scholar 

  • Kagan Y (1960) The kinetics of boiling of a pure liquid. Russ J Phys Chem 34:42–46

    Google Scholar 

  • Kashchiev D (1969) Solution of the non-steady state problem in nucleation kinetics. Surface Sci 14:209–220

    Article  Google Scholar 

  • Landau LD, Lifshitz EM (1958) Statistical physics. Pergamon, London

    Google Scholar 

  • Lifshitz HM, Pitaevskii LP (1981) Physical kinetics. Pergamon Press, Oxford

    Google Scholar 

  • Nishiwaki M, Toramaru A (2019) Inclusion of viscosity into classical homogeneous nucleation theory for water bubbles in silicate melts: reexamination of bubble number density in ascending magmas. J Geophys Res Solid Earth 124(8):8250–8266

    Google Scholar 

  • Schmelzer JW, Baidakov VG (2016) Comment on “Simple improvements to classical bubble nucleation models’’. Phys Rev E 94(2):026801

    Article  Google Scholar 

  • Tanaka KK, Tanaka H, Angèlil R, Diemand J (2016) Reply to “Comment on ‘Simple improvements to classical bubble nucleation models”’. Phys Rev E 94(2):026802

    Google Scholar 

  • Tanaka KK, Tanaka H, Angèlil R, Diemand J (2015) Simple improvements to classical bubble nucleation models. Phys Rev E 92(2):026801

    Article  Google Scholar 

  • Tolman RC (1949) The effect of droplet size on surface tension. J Chem Phys 17(3):333–337

    Article  Google Scholar 

  • Toramaru A (1995) Numerical study of nucleation and growth of bubbles in viscous magmas. J Geophys Res 100:1913–1931

    Article  Google Scholar 

  • Turnbull D, Fisher JC (1949) Rate of nucleation in condensed systems. J Chem Phys 17:71. https://doi.org/10.1063/1.1747055

    Article  Google Scholar 

  • Uhlmann DR (1972) A kinetic treatment of glass formation. J Non-Crystal Solids 7:337–348

    Article  Google Scholar 

  • Zel’dovich YB (1942) Zhur Eksper Teor Fiz 12:525

    Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Atsushi Toramaru .

Rights and permissions

Reprints and permissions

Copyright information

© 2022 The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd.

About this chapter

Check for updates. Verify currency and authenticity via CrossMark

Cite this chapter

Toramaru, A. (2022). Mechanism of Bubble Formation. In: Vesiculation and Crystallization of Magma. Advances in Volcanology. Springer, Singapore. https://doi.org/10.1007/978-981-16-4209-8_3

Download citation

Publish with us

Policies and ethics