Skip to main content

Kinetic Gelation

  • Chapter
  • First Online:
Generalized Statistical Thermodynamics

Abstract

Binary aggregation usually leads to stable populations whose mean size increases indefinitely as aggregation advances, but under special conditions it is possible to observe a phase transition that produces a giant cluster of macroscopic size. In the polymer literature this is known as gelation and is observed experimentally in some polymer systems. Polymerization is perhaps a standard example of aggregation: starting with monomers, polymer molecules of linear or branched structure are formed by joining smaller units together, one pair at a time. The aggregation kernel in this case is determined by the functionality and reactivity of the chemical sites where polymers can join by forming bonds. A kernel that is known to lead to gelation is the product kernel,

$$\displaystyle {} {K}_{i,j} = i j, $$

The classical symptom of gelation is the breakdown of the Smoluchowski equation: within finite time the mean cluster size diverges and the Smoluchowski equation ceases to conserve mass. Gelation has the qualitative features of a phase transition and is often discussed qualitatively in the thermodynamic language of phase equilibrium.

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 149.00
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Hardcover Book
USD 199.99
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.

    This scaling excludes distributions that contain a gel cluster. A necessary though not sufficient condition for Eq. (9.4) is that the distribution must be a single sol phase.

  2. 2.

    Details are given in the appendix, section “Derivations—Power-Law Kernels.”

  3. 3.

    The recursion for the sum kernel is

    $$\displaystyle \begin{aligned} w_i = \frac{i}{2(i-1)} \sum_{j=1}^{i-1} w_{i-j} w_j,\quad \text{(sum)} \end{aligned}$$

    Using a i = iw i the recursion for the product kernel can be written as

    $$\displaystyle \begin{aligned} a_i = \frac{i}{i-1}\sum_{j=1}^{i-1} a_i,\quad \text{(product)} \end{aligned}$$

    Comparing the two recursions we establish the relationship between the cluster functions of the two kernels,

    $$\displaystyle \begin{aligned} w_i^{\text{prod}} = \frac{2^{i-1}}{i} w_i^{\text{sum}}, \end{aligned}$$

    which allows us to obtain w i of the product kernel from that of the sum kernel.

  4. 4.

    See Eq. (8.22) on p. 252.

  5. 5.

    See Smith and Matsoukas (1998).

  6. 6.

    On the other hand, the average mass of the gel cluster is somewhat lower than the position of the maximum, as indicated by the fact that in the vicinity of the gel point ϕ gel from simulation lies below the theoretical value in Fig. 9.4.

  7. 7.

    Compare Flory’s Fig. 2 in Flory (1941) to our Fig. 9.4.

  8. 8.

    Compare Stockmayer’s Fig 2 in Stockmayer (1943) to ϕ gel in our Fig. 9.7.

  9. 9.

    Stockmayer uses w i for this number but here we use a i to avoid confusion with the cluster function w i = a ii!

  10. 10.

    To economize in notation we adopt Stockmayer’s Ω for the microcanonical weight, even though we have reserved it for the partition function. We do this knowing that in the thermodynamic limit their logarithms are equal.

  11. 11.

    In the absence of cycles, it is easy to show that if we move a monomer to a different position on the chain the number of free sites (that is, sites that are bonded to other monomers) does not change. This means that the number of free sites is the same for all polymers with the same number of monomers, regardless of structure. We may then establish this number by looking at a linear chain: the i − 2 monomers in the interior of the chain have f − 2 free sites and the two monomers at the ends of the chain have f − 1 each. The total number is (i − 2)(f − 2) + 2(f − 1).

  12. 12.

    Mayer and Mayer (1940).

  13. 13.

    Wolfram Research, Inc., Mathematica, Version 11.3, Champaign, IL (2018).

References

Download references

Author information

Authors and Affiliations

Authors

Appendix: Derivations

Appendix: Derivations

9.1.1 Derivations: Power-Law Kernels

Here we derive the parameters β and q for the power-law kernel

$$\displaystyle \begin{aligned} {K}_{i,j} = (i\, j)^{\nu/2} . \end{aligned} $$
(9.70)

This includes the product kernel as a special case for ν = 1. For distributions without a gel phase, the mean kernel in distribution n is

$$\displaystyle \begin{aligned} {\bar K}(\mathbf{n}) = \left(\frac{M}{N}\right)^\nu . \end{aligned} $$
([9.58])

The partition function is

$$\displaystyle \begin{aligned} \Omega_{M,N} = \binom{M-1}{N-1} \prod_{g=0}^{M-N-1} \left(\frac{M}{M-g}\right)^\nu = \binom{M-1}{N-1} \left(\frac{M}{M}\frac{M}{M-1} \cdots\frac{M}{N+1}\right)^\nu, \end{aligned}$$

which condenses to

$$\displaystyle \begin{aligned} \Omega_{M,N} = \binom{M-1}{N-1} \left(\frac{M!}{N!}M^{M-N}\right)^\nu . \end{aligned} $$
(9.71)

We obtain β and q in finite-difference form. Starting with β we have

$$\displaystyle \begin{aligned} \beta = \log \frac{\Omega_{M+1,N}}{\Omega_{M,N}} = \nu (M-N) \log \left(\frac{M+1}{M}\right) - \log \left(\frac{M-N+1}{M}\right) . \end{aligned} $$
(9.72)

Using

$$\displaystyle \begin{aligned} \log \left(\frac{M+1}{M}\right)^M\to 1, \end{aligned}$$

we obtain the limiting form of β in the thermodynamic limit:

$$\displaystyle \begin{aligned} \beta = \nu\frac{M-N}{N} - \log \frac{M-N}{N} , \end{aligned}$$

or

(9.73)

For q we find

$$\displaystyle \begin{aligned} q = \frac{\Omega(M,N+1)}{\Omega(M,N)} = \left(1+\frac{1}{N}\right) \left(1-\frac{N}{M}\right) \left(\frac{N+1}{M}\right)^{\nu-1} . \end{aligned} $$
(9.74)

In the thermodynamic limit this becomes

$$\displaystyle \begin{aligned} q = \left(1-\frac{N}{M}\right)\left(\frac{N}{M}\right)^{\nu-1} , \end{aligned}$$

or

(9.75)

With ν = 0, 1, or 2, Eqs. (9.71), (9.73), and (9.75) revert to those for the constant kernel, sum kernel, or product kernel, respectively.

9.1.2 I Locus of Gel Points

The system is unstable if q(θ) contains a branch where dq < 0. The onset of instability is defined by the condition,

$$\displaystyle \begin{aligned} \frac{d q(\theta)}{d\theta} = 0 = (1-\theta )^{\nu -2} (1-\theta\nu) , \end{aligned} $$
(9.76)

whose solution is

$$\displaystyle \begin{aligned} \theta^* = 1/\nu . \end{aligned}$$

Since θ must be between 0 and 1, instability may occur only if ν > 1. The condition that defines the locus of gel points is

(9.77)

The case ν = 1 represents a singular case: the system forms a gel phase but the gel point is at θ = 1, which corresponds to \({\bar x}=\infty \) and is not observed under finite \({\bar x}\).

9.1.3 I Some Useful Power Series

Here are some useful results involving infinite series that are appear in the solution of the sum and the product kernel.

$$\displaystyle \begin{aligned} & S_0 = \sum_{i=1}^\infty \frac{ i ^{ i } } {i!} e^{-i (a-\log a)} = \frac{a}{1-a} \end{aligned} $$
(9.78)
$$\displaystyle \begin{aligned} & S_1 = \sum_{i=1}^\infty \frac{ i ^{ i-1 } } {i!} e^{-i (a-\log a)} = a \end{aligned} $$
(9.79)
$$\displaystyle \begin{aligned} & S_2 = \sum_{i=1}^\infty \frac{ i ^{ i-2 } } {i!} e^{-i (a-\log a)} = a (1 - a/2) \end{aligned} $$
(9.80)

These can be also expressed in equivalent form as

$$\displaystyle \begin{aligned} & \sum_{i=1}^\infty \frac{(i a )^{ i } } {i!} e^{-i a} = \frac{a}{1-a} \end{aligned} $$
(9.81)
$$\displaystyle \begin{aligned} & \sum_{i=1}^\infty \frac{(i a )^{ i - 1} } {i!} e^{-i a} = 1 \end{aligned} $$
(9.82)
$$\displaystyle \begin{aligned} & \sum_{i=1}^\infty \frac{(i a )^{ i - 2} } {i!} e^{-i a} = \frac{1-a/2} {a} \end{aligned} $$
(9.83)

We outline the derivation of these results. We start with the power seriesFootnote 13

$$\displaystyle \begin{aligned} \sum_{i=1}^\infty \frac{i^{i-1}}{i!} e^{-b i} = a, \end{aligned} $$
(9.84)

where a is the solution of

$$\displaystyle \begin{aligned} e^{-b} = a e^{-a} . \end{aligned} $$
(9.85)

Taking the log of this expression we have

$$\displaystyle \begin{aligned} b = a - \log a, \end{aligned} $$
(9.86)

which proves Eq. (9.79). We define

$$\displaystyle \begin{aligned} C_k(i) = \frac{i^{i-k}}{i!}, \end{aligned} $$
(9.87)

and express the summations in Eqs. (9.78)–(9.80) as

$$\displaystyle \begin{aligned} S_k(b) = \sum_{i=1^\infty} C_k(i) e^{-b i} . \end{aligned} $$
(9.88)

We now have

$$\displaystyle \begin{gathered} S_0(b) = -\frac{\partial S_1}{\partial b} \end{gathered} $$
(9.89)
$$\displaystyle \begin{gathered} S_1(b) = a(b), \end{gathered} $$
(9.90)
$$\displaystyle \begin{gathered} S_2(b) = -\int S_1 d b \end{gathered} $$
(9.91)

with the relationship a = a(b) from Eq. (9.86). By differentiation and integration of S 1(b) with respect to b we obtain Eqs. (9.78) and (9.80), respectively. Equations (9.81)–(9.83) follow by straightforward manipulation.

Rights and permissions

Reprints and permissions

Copyright information

© 2018 Springer Nature Switzerland AG

About this chapter

Check for updates. Verify currency and authenticity via CrossMark

Cite this chapter

Matsoukas, T. (2018). Kinetic Gelation. In: Generalized Statistical Thermodynamics. Understanding Complex Systems. Springer, Cham. https://doi.org/10.1007/978-3-030-04149-6_9

Download citation

Publish with us

Policies and ethics