some-topological-aspects-of-fluid-dynamics
some-topological-aspects-of-fluid-dynamics
230
Contents
1 Introduction 3
2 Historical background 4
https://doi.org/10.1017/jfm.2020.230 Published online by Cambridge University Press
5 Frozen-in fields 16
5.1 Frozen-in scalar fields . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
5.2 Frozen-in vector fields; helicity invariance . . . . . . . . . . . . . . . . . 17
5.3 Helicity an invariant of the Euler equations . . . . . . . . . . . . . . . . 18
5.3.1 The Lie derivative . . . . . . . . . . . . . . . . . . . . . . . . . 19
6 Dynamo mechanisms 19
6.1 Turbulent line stretching . . . . . . . . . . . . . . . . . . . . . . . . . . 19
6.1.1 Cranking and helical distortion . . . . . . . . . . . . . . . . . . . 20
6.1.2 Flux-tube distortion by homogeneous turbulence . . . . . . . . . 21
6.2 The slow dynamo . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
6.2.1 The possible growth of small-scale modes . . . . . . . . . . . . . 23
6.2.2 Exponentially growing large-scale force-free modes . . . . . . . . 23
6.2.3 Weak turbulence and the link with helicity . . . . . . . . . . . . 23
6.2.4 The turbulent diffusivity . . . . . . . . . . . . . . . . . . . . . . 24
6.3 The fast dynamo . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
6.3.1 The stretch–twist–fold scenario . . . . . . . . . . . . . . . . . . 25
6.3.2 Curvature, torsion, twist and writhe . . . . . . . . . . . . . . . . 26
6.4 Helicity generated by magnetostrophic turbulence . . . . . . . . . . . . . 27
6.4.1 Up–down symmetry breaking and the ‘αω-dynamo’ . . . . . . . 28
7 Analogies 29
7.1 The B-ω analogy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
7.2 The B-u analogy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
7.3 Flux expulsion and analogous homogenisation . . . . . . . . . . . . . . . 30
8 Magnetic relaxation 32
8.1 The Arnold inequality . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
8.1.1 Energy bound for non-trivial linkage; minimum crossing number . 33
https://doi.org/10.1017/jfm.2020.230 Published online by Cambridge University Press
9 Stability 38
9.1 Stability of magnetostatic equilibria . . . . . . . . . . . . . . . . . . . . 38
9.2 Instability of analogous Euler flows . . . . . . . . . . . . . . . . . . . . 40
9.3 Two-dimensional cylindrical equilibria . . . . . . . . . . . . . . . . . . . 40
9.3.1 Arnold’s assertion and its refutation . . . . . . . . . . . . . . . . 41
9.3.2 Rayleigh’s criterion, as anticipated by Maxwell . . . . . . . . . . 41
9.3.3 Three-dimensional instability of Euler flows . . . . . . . . . . . . 42
9.4 Kelvin modes and transient growth . . . . . . . . . . . . . . . . . . . . . 42
914 P1-2
Some topological aspects of fluid dynamics
10 Knotted flux tubes 43
10.1 Knot helicity, writhe and twist . . . . . . . . . . . . . . . . . . . . . . . 44
10.2 The energy spectrum of knots and links . . . . . . . . . . . . . . . . . . 44
10.3 The analogue Euler knots . . . . . . . . . . . . . . . . . . . . . . . . . . 45
10.4 Tight knots . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
10.5 Experimental realisation of vortex knots . . . . . . . . . . . . . . . . . . 47
References 53
1. Introduction
I welcome this opportunity to write a Perspectives article for JFM, and I thank the Editors
for their invitation to do so. One dictionary definition of ‘perspective’ is ‘a particular
attitude towards or way of regarding something; a point of view’. This gives me freedom
to express my personal opinions throughout the article, and to adopt a more informal style
than is perhaps usual for JFM.
Insofar as fluid dynamics is concerned with continuous deformation induced by flow,
there is a natural symbiosis with topology which is largely concerned with properties
of systems that remain invariant under continuous deformation. I propose to provide a
necessarily superficial survey of a range of topics, all of which have some topological
aspect, in which I have been personally involved at some stage over the last 60 years. Some
of these topics involve flow at low Reynolds numbers, where viscous effects dominate; and
some at high Reynolds numbers where viscous effects are negligible nearly everywhere.
A particular concern in any topological approach is to identify the location and structure
of singularities in a flow field, and the manner in which such singularities can be resolved
(see § 3). A further concern is to identify flow properties that do indeed remain invariant,
and to identify circumstances in which singularities can appear and topological jumps can
https://doi.org/10.1017/jfm.2020.230 Published online by Cambridge University Press
occur; vortex reconnection is perhaps the best known circumstance of this kind, and my
discussion will build up to a brief consideration of this problem and the implications for
turbulence in § 11.
Magnetohydrodynamics plays an important part here in that, in an ideal conducting
fluid, the magnetic field is ‘frozen in’, i.e. transported with the fluid (§ 5). Analogies
with vortex dynamics and with steady Euler flows can be powerful in their implications,
but must be treated with caution (§ 7). Topological properties are particularly relevant in
both fast- and slow-dynamo theory (§ 6) and in the theory of magnetic relaxation (§ 8)
which raises issues of stability (§ 9). This leads naturally to questions concerning the
existence and structure of knotted flux tubes, and of field discontinuities that are inevitably
encountered (§ 10).
My research in fluid dynamics started in 1958 under the supervision of George Batchelor
FRS, whose centenary will be celebrated by a special IUTAM Symposium to be held
in Cambridge, 15–18 March 2020.1 In 1958, Batchelor was, at 38 years old, a world
authority on turbulence, and he had founded this Journal just two years earlier (for details
1 Now postponed because of the COVID-19 pandemic to an online symposium, 28–31 March 2021.
914 P1-3
H.K. Moffatt
concerning this great achievement, see Moffatt (2017)). He was also at that time engaging
with the authorities of Cambridge University in creating the Department of Applied
Mathematics and Theoretical Physics (DAMTP), officially established in 1959. It was
natural that I should undertake research in some aspect of turbulence, and I settled on
Magnetohydrodynamic Turbulence (the title of my PhD thesis), magnetohydrodynamics
being then at a very exciting stage of development following publication of the Interscience
texts of Spitzer (1956) and Cowling (1957). Batchelor was a superb research adviser,
encouraging and critical at the same time, and unfailing in the good advice he gave at
all stages of my early faltering attempts to grapple with ‘the problem of turbulence’.
2. Historical background
2.1. Helmholtz’ laws
My story starts with the seminal paper of Helmholtz (1858), who stated his three laws
of vortex motion for flow of an ‘ideal fluid’ in a bounded domain, laws which may be
paraphrased as follows: (i) a vortex tube has constant circulation (i.e. flux of vorticity)
along its length; (ii) a vortex tube must either be closed on itself or terminate on the fluid
boundary; and (iii) vortex lines are transported with (or ‘frozen in’) the flow. This paper by
Helmholtz was translated from German into English by Tait (1867), and came immediately
to the attention of William Thomson (later Lord Kelvin) who recognised the particular
significance of Helmholtz’s law (iii), and immediately proposed his ‘vortex atom’ theory
(Thomson 1867 – see below).
The first law (i) is merely a way of saying that ∇ · ω = 0, which of course follows
immediately from the definition of vorticity: ω = ∇ ∧ u, where u(x, t) is the velocity field.
We shall use the symbol Γ for the circulation of a vortex tube. The term ‘vortex filament’
may be used to describe a vortex tube of infinitesimal cross-section.
The statement of the second law (ii) is false, as now widely recognised, because in any
flow that exhibits the (generic) phenomenon of chaos (see § 4 below), a vortex line in
any chaotic sub-domain of the flow wanders indefinitely without ever closing on itself.
Saffman (1993) has maintained that the statement (ii) can be rescued by simply replacing
‘vortex lines’ by ‘vortex tubes’. In § 1.4 of his well-known book on Vortex Dynamics he
https://doi.org/10.1017/jfm.2020.230 Published online by Cambridge University Press
wrote ‘If the vorticity field is compact, the tubes must be closed or begin and end on
boundaries’. But this too is false; for in any chaotic sub-domain, any two neighbouring
vortex lines diverge exponentially, and the cross-section of any vortex tube becomes
increasingly flattened and distorted along its length; it will in general partially overlap
itself, and does so repeatedly in these circumstances, but cannot surely be regarded as
‘closed’. (I made this point in my review of Saffman’s book (Moffatt 1994), and, following
its publication, enjoyed an extensive correspondence with him about chaotic vector fields.)
(c)
Figure 1. (a) First page of James Clerk Maxwell’s letter to Peter Guthrie Tait, 13 November 1867; (b,c) Tait’s
frequent method of reply to Maxwell’s letters. (Reproduced by kind permission of the Syndics of Cambridge
University Library.)
(reproduced from the original in figure 1a), Maxwell, with a degree of gentle scepticism,
expresses his views concerning Thomson’s ‘worbles’: he talks of ‘the interpretation
Thomson has set himself to spin the chains of destiny out of a fluid plenum . . .’ and
adds ‘I saw you had put your calculus in it too. May you both prosper and disentangle
your formulæ in proportion as you entangle your worbles’. (This was the beginning of
an extended correspondence between Maxwell and Tait, who had been close friends
ever since their schooldays at the Edinburgh Academy; Tait would frequently reply to
https://doi.org/10.1017/jfm.2020.230 Published online by Cambridge University Press
914 P1-6
Some topological aspects of fluid dynamics
(a) (b) (c)
3 3 3
2 2 2
1 1 1
0 0 0
–1 –1 –1
–2 –2 –2
–3 –3 –3
–3 –2 –1 0 1 2 3 –3 –2 –1 0 1 2 3 –3 –2 –1 0 1 2 3
Figure 2. Figure illustrating the merging of two stagnation points (an extremum and a saddle) as t increases
through zero for the streamfunction ψ(x, y, t) = y2 − x3 − 3xt; (a) t = −1, (b) t = 0, (c) t = +1; the cusped
streamline exists only instantaneously at time t = 0.
0.5
0.4
0.3
0.2
0.1
0
–1.0 –0.5 0 0.5 1.0
Figure 3. Stokes flow described by the streamfunction ψ = y2 ( y − kx), with no slip on the boundary y = 0.
√
x = + −t (a saddle). The separation of these points decreases at speed ∼(−t)−1/2 ,
and they merge (at infinite speed!) at t = 0. Note the cusped structure of the critical
streamline ψ1 = 0 at t = 0. This is an instantaneous topological transition, of a type that
regularly occurs in the evolution of meteorological maps. We note further that ψ1 trivially
satisfies the unsteady non-dimensionalised Stokes equation ∂∇ 2 ψ1 /∂t = ∇ 4 ψ1 , so that,
since the nonlinear inertia force is negligible near the stagnation points, this transition is
https://doi.org/10.1017/jfm.2020.230 Published online by Cambridge University Press
as had been suggested by Rayleigh (1920, p. 18) and pursued by Dean & Montagnon
(1949), was to assume a streamfunction of the form
ψ(r, θ) = rλ f (θ), (3.1)
in plane polar coordinates, to substitute this in the biharmonic equation ∇ 4 ψ = 0
governing the Stokes flow near the corner, and then to seek to determine λ by satisfying
the no-slip conditions on the bounding planes. If it is supposed that ψ is symmetric
approximately θ = 0, this leads to the equation
sin 2μα = −μ sin 2α where μ = λ − 1. (3.2)
The novel property of this equation is that all non-zero solutions for μ are complex if
2α 146.3◦ . This much had been discovered by Dean & Montagnon (1949), but the
fact that this implies infinite oscillations as r → 0 was not recognised by these authors.
I found it difficult to believe this myself, and I took the prediction to George Batchelor,
expecting him to say there must be a mistake somewhere. To my great relief, he said
‘Yes, I can believe that’, and gave me every encouragement to write my paper on this
topic (Moffatt 1964); he obviously approved of my efforts to extract physical meaning
from such a curious mathematical result! The function f (θ) in (3.1) is also complex, but
since the Stokes problem is linear, we may simply redefine ψ as Re [rλ f (θ)]. The resulting
streamfunction exhibits a geometric sequence of counter-rotating eddies, as illustrated for
https://doi.org/10.1017/jfm.2020.230 Published online by Cambridge University Press
the case α = π/18 in figure 4(a). The important thing about this flow is that it exhibits the
phenomenon of flow separation and reattachment where the dividing streamlines ψ = 0
meet the boundaries θ = ±α. Near these points, the flow is just as described in figure 3,
with k > 0 for separation and k < 0 for reattachment. Separation had previously been
thought of as a high-Reynolds-number phenomenon; but here it was also evident, and
quite dramatically so, at low Reynolds number also.
The flow shown in figure 4(a) was realised experimentally by Taneda (1979), who
observed the first two eddies in a sequence driven by rotation of a cylinder far from the
corner (figure 4b). A third eddy could also be dimly discerned, although the velocity in
it was extremely small. The theory does indeed imply a rapid decrease in flow intensity
from one eddy to the next as the corner is approached – by a factor of approximately 400
when α = π/18. If the first eddy has a circulation time of say 10 s, then the second will
have a circulation time ∼1 h, the third ∼16 days, and the fourth ∼17 years; to observe
such eddies demands patience! Indeed, the fluid is virtually stagnant after the third eddy
in the sequence, whatever the remote stirring mechanism may be; and yet, because λ is not
an integer, high derivatives of the velocity are infinite at r = 0 for nearly all values of the
angle α!
914 P1-8
Some topological aspects of fluid dynamics
–0.2×10–9
(a) (b) –0.529×10–9
–0.3×10–6 0 V3
–0.15×10–5
Axis of –0.261×10–5
V2
curvature y 0
A
–0.5 ×10–4
β
–0.25 ×10–3
0.18
V1
κ−1 0
–0.425 ×10–3
D
o 2α –3
×10
0.1
β 0–
2
0 ×1 Main vortex
1
.2
–
×10
B 0.1 –1
0.3 ×10
0.18
Figure 5. The curved duct configuration of Collins & Dennis (1976). When the flow is pressure driven, eddies
form at A and B if β > 40.4◦ , and at O if 71.9◦ < 2α < 159.1◦ . When the flow is driven by rotation of the
boundary AB about the axis of curvature, eddies do not form at A and B, but they do form at O if 35.0◦ <
2α < 159.1◦ (after Collins & Dennis 1976).
3.3. Universality
The beauty of the corner flow solution ψ ∼ Re rλ f (θ) lies in what may be described
as the ‘universality’ of the phenomenon that it describes. First, although this appears to
be a low-Reynolds-number phenomenon, this form of ψ actually provides an asymptotic
914 P1-9
H.K. Moffatt
solution of the Navier–Stokes equations for arbitrary ‘driving Reynolds number’ Re (i.e.
based on the driving mechanism far from the corner). This is because both the local length
scale and the flow velocity tend to zero as r → 0. Thus the Stokes separation phenomenon
is universal for arbitrary Re and arbitrary two-dimensional flow near a corner, provided that
the angle of the corner is 146.3◦ . The location of the first separation point depends on the
remote forcing mechanism; moreover, this location will be Reynolds-number dependent in
a manner that still calls for detailed investigation.
Second, even if the corner is not sharp (and no corner is perfectly sharp in reality), the
flow will still in general in the low-Re regime exhibit a sequence of counter-rotating eddies,
but the number of these will be finite; indeed if flow is driven by a rotating cylinder placed
in a converging channel, it may be expected to exhibit a similar eddy sequence. If there
is a weak superposed flow through the channel, then the eddies are attached alternately to
the walls of the channel, allowing the flow to pass between them.
A similar phenomenon occurs at a cusped corner, e.g. in steady shear flow over a
cylinder that sits on a plane boundary: the flow separates and a sequence of eddies appears
in the cusp regions, both upstream and downstream because (at low Re) the flow exhibits
symmetry about the diameter of the cylinder through its point of contact with the plane.
If there is a small gap between the cylinder and the plane, then there is a small leakage
of fluid through the gap, and the eddies, again finite in number, are in this case attached
alternately to the cylinder and the plane (Jeffrey & Sherwood 1980).
(b) (c)
ace 2γ
surf
Free Circle of
curvature
d radius R
Vortex dipole
α
Figure 6. (a) A cusp at the free surface of a viscous liquid induced by sub-surface counter-rotating cylinders
(the cylinder on the left rotates clockwise, that on the right anti-clockwise); the black streak entering the fluid
from the cusp marks a thin sheet of air that enters the bell-shaped bubble which is held stationary in the
downward flow; (b) flow modelled by a vortex dipole of strength α at depth d below the position of the free
surface when undisturbed; the cusp appears at depth 2d/3; (c) local situation near the stagnation point on the
plane of symmetry (adapted from Jeong & Moffatt 1992).
playing field’ as between viscosity and surface tension (i.e. C = 1) then (3.4) gives the
extraordinary result
https://doi.org/10.1017/jfm.2020.230 Published online by Cambridge University Press
Figure 7. As in figure 6, but here the cylinders are close to each other, and only partially submerged; (a) the
viscous fluid is drawn up in a layer on each cylinder and the layers interact as the fluid passes down through
the gap, forming a cusp; the free surface can be seen on the right of the photo; (b) blow-up of the cusp region
showing how air is drawn through the cusp in a very thin sheet forming a ‘tricuspidal’ bubble from which
smaller bubbles of air are ejected into the fluid. (Photographs taken by author in 1992, but not previously
published.)
U i
Figure 8. (a) Hypothetical (but unrealistic) flow near the contact line when a flat plate is drawn into a viscous
fluid with velocity U; (b) the ‘half-cusp’ between the free surface and the plate that must occur near the contact
line due to the downward drag on the fluid.
4. Lagrangian chaos
4.1. ABC flow
The flows considered so far have been regular in the sense that the streamlines and/or
particle paths are either closed curves or curves confined to a family of surfaces. The
generic structure of steady flows in three dimensions does not satisfy either of these
constraints; in general, there exist subdomains within the fluid in which the streamlines
are space filling: they wander in such a way as to come arbitrarily near any point of the
subdomain if followed far enough. Such flows exhibit what is described as ‘Lagrangian
chaos’. The behaviour occurs also in unsteady two-dimensional flows, as exemplified by
the ‘blinking vortex’ model of Aref (1984).
Chaos in fluid flows was a subject that sprang to life with the work of Arnold (1965a)
and Hénon (1966), who studied what came to be known as the ABC flow,
u(x) = (B cos ky + C sin kz, C cos kz + A sin kx, A cos kx + B sin ky), (4.1)
914 P1-13
H.K. Moffatt
(a) (b)
Figure 9. Sample Poincaré sections for the ABC flow; (a) A2 = 1, B2 = 2/3, C2 = 1/3, showing islands of
regularity in a sea of chaos; (b) the contrasting situation when A2 = 1, B2 = 1, C2 = 1; the region of chaos is
very much reduced. (From Hénon 1966; Dombre et al. 1986.)
which satisfies the Beltrami condition ω(x) = ku(x). Any incompressible flow uB
satisfying this condition (with k constant) also satisfies the condition ∇ 2 uB ≡ −∇ × ωB =
−k2 uB , and therefore satisfies the Navier–Stokes equation (linear for such flows),
in English translation since 1994). Thus the streamline structure remains constant under
Navier–Stokes evolution in this very special situation.
The flow (4.1), being periodic in x, y and z, can be treated as a flow on the three-torus
T3 , a description that may be less than helpful for those who prefer to remain firmly in the
Euclidean space R3 , in which the flow can actually occur. Nevertheless, it is on T 3 that the
streamlines of the flow are chaotic. This chaos has been studied in some detail by Dombre
et al. (1986) who summarise their results with the statement ‘In general, there is a set of
https://doi.org/10.1017/jfm.2020.230 Published online by Cambridge University Press
closed (on the torus T3 ) helical streamlines, each of which is surrounded by a finite region
of Kolmogorov–Arnold–Moser invariant surfaces. For certain values of the parameters
strong resonances occur which disrupt the surfaces. The remaining space is occupied
by chaotic particle paths: here stagnation points may occur and, when they do, they are
connected by a web of heteroclinic streamlines’. A typical Poincaré section is reproduced
in figure 9(a) for the particular case A2 = 1, B2 = 2/3, C2 = 1/3, showing ‘islands of
regularity’ within a sea of chaos which extends over roughly half the fluid domain; a
single streamline here provides the scatter of points in the chaotic region. Generally, it
appears that the region of chaos decreases in extent as the parameters A, B and C approach
equality, although a modest extent of chaos survives in the limiting situation, as shown in
figure 9(b).
(a) (b)
Figure 10. (a) Typical streamline of a flow of the form ((4.3), (4.4a–c)) which indicates two strong vortices;
(b) Poincaré section for the same streamline by a plane perpendicular to the vortices, showing points where it
has crossed the plane of section 40 000 times. (From Bajer & Moffatt 1990.)
& Quake 2005). A different type of Lagrangian chaos can exist in such systems.
Figure 10(a) shows a typical streamline for a steady Stokes flow in a sphere, and
figure 10(b) an associated Poincaré section for the same streamline when it is continued for
a very long time. In figure 10(a), the streamline appears to lie on a surface; but the Poincaré
section shows that this is not in fact the case: the ‘surface’ shifts by random small amounts
when it nearly returns on itself – a phenomenon described as ‘transadiabatic drift’ (Bajer
& Moffatt 1990).
The particular Stokes flow with streamlines as in figure 10 is one of a class of steady
flows consisting of three ingredients, each of which is an incompressible Stokes flow
confined to the sphere r < 1
u(x) = U(x) + V (x) + W (x), (4.3)
where
U(x) = a(1 − 2r2 ) + (a · x)x, V (x) = Ω × x, W (x) = (λyz, μzx, νxy), (4.4a–c)
and where λ + μ + ν = 0 (so that W · n = 0 on r = 1). Here, U(x) is the same as the
https://doi.org/10.1017/jfm.2020.230 Published online by Cambridge University Press
flow inside a Hill’s spherical vortex, axisymmetric about the vector a, V (x) is a rigid body
rotation with angular velocity Ω and W (x) is a combination of ‘twist ingredients’; this
type of flow was originally devised to represent the ‘stretch–twist–fold’ process, believed
to be fundamental for dynamo theory (see § 6.3 below). Each such flow u(x) inside the
sphere r = 1 has to be driven by a non-zero tangential velocity on the surface r = 1 and
the associated tangential stress; thus energy is pumped into the sphere from the surface
and dissipated internally by viscosity. If the amplitude of the flow is normalised (e.g.
by setting λ2 + μ2 + ν 2 = 1), there remains a seven-parameter family of flows of this
kind, all quadratic functions of the space coordinates, all Stokes flows in a sphere, and
all exhibiting some degree of chaos except in limiting situations, as described in Bajer &
Moffatt (1990). The flow shown in figure 10 looks as if the streamline lies on a surface
around two vortices; but in fact when this streamline (or equivalently particle path) is
continued for a long time, the Poincaré section by a plane perpendicular to the vortices
shows a high degree of chaos in the flow.
A similar situation arises when a small drop, kept spherical by surface tension, is
subjected to a general strain field in the surrounding fluid (Stone, Nadim & Strogatz
914 P1-15
H.K. Moffatt
1991). The Stokes flow inside the drop in this situation is a cubic function of the
cartesian coordinates, and again the particle paths in general exhibit Lagrangian chaos.
An important, indeed defining, property of this type of Lagrangian chaos is that initially
neighbouring particle paths diverge exponentially when averaged over a long time (i.e. the
Lyapunov exponent is positive), at least until the separation is comparable with the scale
of the drop. A small blob of dye in such a flow is stretched into a thin highly convoluted
sheet as time progresses. The behaviour is conducive to strong mixing, important when
homogeneity within the droplet is the objective.
5. Frozen-in fields
The topological aspect of fluid mechanics is most prominent in consideration of properties,
whether scalar or vector or even higher-order tensor, that are transported with the flow. For
example if a dye is used to colour a subdomain DL of the fluid, and if molecular diffusion
is neglected, then this coloured region is obviously transported with the flow, i.e. DL is
a Lagrangian subdomain. We say that the dye is ‘frozen in the fluid’, or simply that it is
a ‘frozen-in’ field. As recognised by Helmholtz, vorticity in an ideal fluid is a frozen-in
vector field, the vortex lines being transported with the flow and the circulation round any
material (Lagrangian) circuit being conserved. These are the two most familiar examples
of frozen-in fields, which we now consider in more detail.
the fluid’. For a localised ‘blob’ of dye, the surfaces θ = const. are closed, and since they
move with the fluid, the volume of fluid within each such surface is constant. In a chaotic
or turbulent flow, the area of each surface element increases exponentially on average. The
volume δV between any two neighbouring surfaces labelled θ and θ + δθ is conserved, so
their separation decreases exponentially on average. This implies an exponential increase
in |∇θ| on average.
This phenomenon was first recognised, in the context of homogeneous turbulence, by
Batchelor (1952). Batchelor supposed that θ(x, t) was, like u(x, t), a stationary random
function of x, and that it is measured relative to its mean, so that θ = 0. In these
circumstances, (5.1) implies that
d θ2
= −2κ G 2 , (5.2)
dt
where G = ∇θ , so that θ 2 decays to zero as a result of molecular diffusivity κ. (The
angular brackets here may be interpreted as a space average.) At the same time, if κ
is sufficiently small, G2 certainly increases exponentially for so long as diffusion is
914 P1-16
Some topological aspects of fluid dynamics
negligible, by the mechanism indicated above. This increase is associated with transfer
of the spectrum of θ 2 to progressively higher values of wavenumber k until diffusion is
no longer negligible, and statistical balance between convection and diffusion is attained.
From this point on, G2 must decay in tandem with the persistent decay of θ 2 to zero.
The conclusion that G2 increases exponentially to a maximum before decaying to zero
is one that still calls for numerical investigation, which should at the same time seek
to determine the dependence of the maximum attained by G2 on the turbulent Péclet
number Pe = u0 0 /κ, where now u0 and 0 are velocity and length scales characterising
the energy-containing eddies of the turbulence. I am not aware of any such study,
although there have been many numerical investigations of the corresponding even more
challenging problem of a transported vector field, to which I now turn.
+ η∇ 2 A − ∇ϕ,
= u · ∇A (5.6)
Dt
where the tilde ˜· indicates the transpose, and ϕ is an arbitrary gauge field. When η = 0,
combining (5.5) and (5.6) leads without difficulty to the equation
D
(A · B) = (B · ∇)(A · u − ϕ). (5.7)
Dt
Integrating this equation over any Lagrangian volume VL bounded by a ‘magnetic surface’
on which n · B = 0 yields the equation
d
A · B dV = 0, (5.8)
dt VL
with the consequence that the magnetic helicity HM = A · B dV, when integrated over
any volume bounded by a magnetic surface, is invariant. A more limited result of this kind
was first obtained by Woltjer (1958). HM may be positive or negative; it is a pseudo-scalar,
changing sign under change from a right-handed to a left-handed frame of reference.
914 P1-17
H.K. Moffatt
If η =
/ 0, this invariance is broken. For example, if we consider a localised B-field in a
fluid of infinite extent, then (5.8) is replaced by
d
A · B dV = −2η B · ∇ × B dV. (5.9)
dt
If HM = 0 at any instant, then it is evident from this that if B · ∇ × B dV =
/ 0, HM will
not remain zero, but will be generated by the diffusive process.
∂ω
= ∇ × (u × ω) + ν∇ 2 ω. (5.10)
∂t
the integral now being over any Lagrangian volume bounded by a ‘vorticity surface’ on
which ω · n = 0. This helicity is invariant under precisely the same three conditions under
which Kelvin’s classic circulation theorem holds: (i) the fluid is inviscid; (ii) the flow,
https://doi.org/10.1017/jfm.2020.230 Published online by Cambridge University Press
6. Dynamo mechanisms
Dynamo theory is concerned with the generation and maintenance of magnetic fields such
as those that are observed in planets, stars and galaxies. In planets such as the Earth, the
field can be generated in the liquid conducting core; in stars like the Sun, in the ionised gas
of the turbulent convecting zone; and in galaxies like the Milky Way, in the ionised gas
of the interstellar medium. The fact that a magnetic field B is frozen in (in the perfectly
conducting limit η = 0) implies intensification due to field-line stretching, a process that
is particularly effective in turbulent flow as already recognised by Batchelor (1950). This
intensification is, however, to some extent compensated by ohmic diffusion when η = / 0,
and the crucial question is then this: in the battle between the two processes, intensification
vs diffusion, which will prevail over the long term? A signal achievement of turbulence
theory over the past 60 years has been to provide a convincing answer to this key question;
this is that in general intensification will prevail provided the mean helicity of the turbulent
flow is non-zero over sufficiently large subdomains of the fluid region. Some aspects of
this major field of research will be discussed in this section.
due to Orszag (1977). Let x(a, t) be the (random) position of the fluid particle initially at
position a. Then δxi (t) = Dij δaj , where Dij = ∂xi /∂aj is the deformation tensor, satisfying
det Dij = 1, by virtue of incompressibility. It follows that δx2 = Wjk δaj δak , where, in
matrix notation, W = DT D. Since W is real and symmetric, its eigenvalues w1 , w2 , w3
are real (and in general unequal), and w1 w2 w3 = det W = (detD)2 = 1.
Now, since Wjk is a statistical property of the turbulence, here assumed homogeneous
and isotropic, it must also be isotropic, i.e.
Wjk = λ(t)δjk , where λ(t) = 1
3 Wii = 1
3 w1 + w2 + w3 > (w1 w2 w3 )1/3 = 1,
(6.1)
since the arithmetic mean of w1 , w2 , w3 (which are certainly not everywhere equal) is
greater than the geometric mean. It follows that
δx2 = Wjk δaj δak = λ(t)δa2 > δa2 . (6.2)
This argument on its own is not sufficient to show that δx2 must systematically
increase in time. After all, the same argument could be applied to random vibrations of an
914 P1-19
H.K. Moffatt
(a) (b) y
2.0 2 1
1.5 0 –1
1.0 y –2
0.5
1.0 0 4
0.5
z 0
–0.5 2 z
–1.0
–2 0 2
x 0
0 5
–5
x
Figure 11. Distortion of the line initially coincident with the x-axis by two velocity fields; (a) a cranking
distortion shown at time t = 10π, at which stage the blue curve has been cranked through 5 complete
turns about the line y = 1, z = 0; at all times, the curve lies on the surface (shaded) with parametric
equations (x, (1 + 2x2 )−1 (1 − cos t), (1 + 2x2 )−1 sin t); (b) distortion by a helical velocity field of the form
u = (0, r Ω(r), w(r)) in cylindrical polar coordinates (r, θ, z), and with Ω(r) = exp(−0.3r2 ) and w(r) =
exp(−r2 ); the z-component of velocity raises the curve locally to the shape of a gaussian, and the θ -component
simultaneously rotates the central part of the loop of the gaussian about the z-axis; the curves shown are at
times t = π/2 (blue), and t = 3π/2 (red).
elastic medium, for which δx2 is presumably time-periodic. A further property of fluid
turbulence is needed to give systematic increase; this is finite time correlation, i.e. finite
‘memory time’. Thus for example, if we introduce a correlation time tc with the property
that ui (x, t)uj (x, t + τ ) = 0 for τ > tc , then the above argument may be iterated in each
time interval ntc < t < (n + 1)tc , giving the systematic trend that is to be expected on
physical grounds.
the stretching is compensated by ohmic diffusion, and then a statistically steady state, as
envisaged above, seems a possible outcome.
The problem was taken up by Saffman (1963), who, always the iconoclast, challenged
the conclusion of Batchelor, and in fact came to the opposite conclusion that the increasing
importance of diffusion as the scale of the field decreases would lead instead to ultimate
decay to zero of the magnetic energy (just as for the decay of G2 as discussed in
§ 5.1 above). For the flux-tube model introduced above, this ultimate decay is indeed
a possibility if the predominant action of the turbulence is to bring oppositely directed
portions of the tube into close proximity, in which case swift local annihilation will occur.
But there seems no good reason why this should occur everywhere for the flux-tube
evolution as considered above; it does not occur, for example, for either of the distorting
motions shown in figure 11.
u × b − u × b (so that F = 0). The essential thing now is to find an expression for E
in terms of B0 , so that (6.5) may be integrated. In a way, this may be seen as a classic type
of ‘closure problem’.
Now, for ‘given’ u, it is evident that (6.6) establishes a linear relationship between b and
B0 , and so between E = u × b and B0 . This must take the form
Ei = αij B0j + βijk ∂B0j /∂xk + · · · , (6.7)
where the coefficients αij , βijk are (pseudo-tensor) statistical properties of the turbulence,
dependent indirectly on the parameter η. If the turbulence is assumed to be isotropic (i.e.
statistically invariant under rotations) then these pseudo-tensors must also be isotropic, i.e.
αij = α δij and βijk = β ijk , so that then
E = αB0 − β∇ × B0 + · · · , (6.8)
where now α is a pseudo-scalar and β a pure scalar (since the ‘pseudo’ character is taken
up by the pseudo-tensor factor ijk ). Being a pseudo-scalar, α changes sign under parity
transformation (change from a right-handed to a left-handed frame of reference), and can
914 P1-22
Some topological aspects of fluid dynamics
therefore be non-zero only if the turbulence lacks reflection symmetry, i.e. is chiral in
character. This is not inconsistent with the assumed isotropy: turbulence, like a box of
vigorously shaken right-handed screws, can be statistically invariant under rotations while
being non-invariant under reflections. The possibility of a mean emf parallel to the mean
field, as indicated by the first term of (6.8) is what is known as the ‘alpha effect’, and is
what has transformed our whole understanding of the turbulent dynamo process.
and whereH(k, ω) is the ‘helicity spectrum’ of the turbulence, with the property that
u·ω = H(k, ω) dk dω. Thus, α is simply a weighted integral of this helicity
spectrum, which is indeed a pseudo-scalar property of the turbulence.
Two particular properties of the expression (6.13) are worth noting. First, if η = 0, and
provided the integral then converges at ω = 0 (a sufficient condition for this is H(k, ω) =
O(ω1+γ ) with γ > 0), then α = 0 also; i.e. the α-effect, and so the above type of dynamo
instability, requires non-zero diffusivity to be operative. This is quite surprising: after all,
ohmic diffusion is responsible for the decay of a random magnetic field in the absence
of fluid motion. Here it appears that diffusion is also responsible for the generation of
magnetic field through turbulent dynamo action.
Second, the minus sign in (6.13) is to be noted. If the helicity spectrum is positive for all
{k, ω}, then α is negative, and since K in (6.12) must have the same sign as α for instability,
the growing magnetic field must have negative magnetic helicity. More generally, the
helicity of the growing field has the opposite sign from the weighted helicity spectrum
of the turbulence given by the integral in (6.13).
The limiting situation when η is so small that Rm = u0 0 /η 1 is relevant in
astrophysical contexts, and deserves particular attention. It is usual in this situation to
assume on a primitive ‘mixing-length’ basis that the turbulent diffusivity β is of order
u0 0 . On the same basis, one might suppose that α should be of order u0 . But then, the
length scale of maximum growth rate (from (6.12) is of order β/α = O(0 ), and this is
inconsistent with the two-scale assumption L 0 . However, the assumption α = O(0 )
implicitly supposes that the turbulence is maximally helical, in the sense that u · ω =
O(u20 /0 ), a state that is unlikely to be attained in practice. If the helicity is relatively weak
https://doi.org/10.1017/jfm.2020.230 Published online by Cambridge University Press
(i.e. u · ω u20 /0 ), then α β/0 also, and the condition L ∼ β/α 0 survives. As
far as I am aware, this justification for mean-field theory in the limit Rm 1 (Moffatt &
Dormy 2019, § 9.2.1) is relatively new, and deserves to be tested numerically.
0.8
0.6
0.4
0.2
the extremely sparse structure of the magnetic field is already evident. It is hypothesised
that the width of the near-singular sheets apparent in this figure is O(Rm −1/2 ) as Rm →
∞, consistent with the general conjecture of Moffatt & Proctor.
914 P1-25
H.K. Moffatt
(a) (b) (c) (d)
Figure 13. The stretch–twist–fold process, illustrated by the parametric equation (6.14); here stretching,
twisting and folding occur simultaneously as t increases from 0 to 1. A shaded tube encloses the central curve;
the cross-section of this tube decreases so that the volume of the tube remains constant. An inflexion point
occurs at s = 0 at time t = 0.2. (a) t = 0. (b) t = 0.2. (c) t = 0.75. (d) t = 0.95.
(a) (b)
2.0 5
1.0 τ 0
c 1.0 –5
1.0
0.5
0 –10
0.5 0.5
–2 t t
–2
0 0
s 2 0 s 2 0
Figure 14. (a) Curvature c(s, t) and (b) torsion τ (s, t) of the family of curves (6.14) shown in figure 13;
singular behaviour is evident at the inflexion point s = 0 when t = 0.2.
(Maggioni & Ricca 2006). This evolution is represented by the blue curves in figure 13. We
place an untwisted magnetic flux tube of small cross-section and invariant volume around
these curves, as illustrated. A velocity field u = U(x, t) is then effectively known on the
surface S of this tube. We may then define a velocity field u(x, t) everywhere outside the
tube as the unique quasi-static Stokes flow satisfying the boundary conditions u = U(x, t)
on S, and u → 0 at ∞. This then is a velocity field that can effect at least one iteration of
the stretch–twist–fold process; and periodic repetition in each successive unit time interval
might come near to the process envisaged by Vainshtein & Zel’dovich.
https://doi.org/10.1017/jfm.2020.230 Published online by Cambridge University Press
The ‘writhe’ Wr(t) of the curve (6.14) can be computed as a function of t from the
formula
1 (dx × dx ) · (x − x )
Wr(t) = . (6.15)
4π C C |x − x )|3
The result is shown in figure 15(c). This writhe, as might be expected decreases from zero
to −1, the final value reflecting the single (negative) self-crossing of the curve in the limit.
The twist Tw(t) = T (t) + N (t) rises increases from zero to +1, the sum Tw(t) + Wr(t)
being constant (here zero), a result in differential geometry familiar to anyone who has
sought to straighten a coiled garden hose!
inductive processes in the Earth’s liquid metallic core are responsible for the generation
and evolution of the geomagnetic field. Here convection, either thermal or compositional,
is driven by the release of buoyant fluid from the ‘mushy zone’ at the boundary of the
slowly solidifying inner core. The dynamics in the outer liquid core is dominated on the
one hand by Coriolis forces due to the global rotation Ω, and on the other by the Lorentz
force associated with the strong toroidal magnetic field B0 that is generated as part of the
dynamo process by differential rotation acting on the poloidal field. Any buoyant parcel
has to navigate upwards under the influence of these forces.
An attempt to understand such upward migration was initiated by Moffatt & Loper
(1994); this attempt led to the numerical investigation of St. Pierre (1996), who found
that such a parcel is subject to a ‘slicing’ instability, the slices being in planes parallel to
both Ω and B0 . This discovery in turn stimulated a reformulation of the problem without
any prior assumption concerning the detailed structure of the buoyancy field θ(x, t),
but merely supposing that this field is statistically homogeneous with ‘given’ spectral
properties (Moffatt 2008). (This can be considered as one step on from the ‘kinematic’
approach to turbulent dynamo theory, which assumes that it is the velocity field u(x, t)
that is statistically prescribed.)
914 P1-27
H.K. Moffatt
The dominant nonlinearity in this approach turns out to be the convective term u · ∇θ
in the advection–diffusion equation
will increase exponentially (bearing in mind the exponential stretching associated with
the term ∇ × (u × B)). As Saffman (1963) pointed out, this fails to recognise that field
stretching may be associated with decrease of scale, and so of accelerated ohmic diffusion.
Nevertheless, Batchelor’s conclusion may be correct, and direct numerical simulation
(DNS) studies do tend to support it. However, DNS studies also show dynamo growth
of small-scale magnetic field even when Pm 1, whereas in this range of Pm , Batchelor
argued that the field would necessarily decay.
So there is a distinct uncertainty here, which calls for more extensive DNS studies. If
Pm < 1, how exactly is it that (in reflectionally symmetric turbulence) the field energy
can grow despite the relatively strong ohmic diffusion? And is there a critical value
of Pm below which the field energy does decay? Although 70 years have elapsed since
Batchelor’s seminal paper, as far as I know these questions still remain unanswered.
The analogy between (7.1) and (7.2) does have one legitimate consequence, namely that,
just as magnetic helicity is conserved in a perfectly conducting fluid (η = 0), so kinetic
helicity is conserved in an ideal inviscid fluid (ν = 0), and indeed, as already mentioned,
it was through exploitation of this analogy that I stumbled on the invariance of helicity
914 P1-29
H.K. Moffatt
in 1969. This is legitimate because the proof of invariance of helicity makes no appeal to
the constraint ω = ∇ × u; it requires only that the vorticity field ω(x, t) satisfy the frozen
field (7.2).
The constraint ω = ∇ × u which does not apply to the field B(x, t) does nevertheless
indicate a serious weakness in the analogy, which can lead to misleading conclusions. For
one thing, there is a freedom in the choice of B(x, 0) which is not available to ω(x, 0),
and this means that there can exist modes of growth of B2 that are not available to
ω2 . This becomes of crucial importance when we consider chiral turbulence (lacking
reflection symmetry) as in § 6.2 above, and it was here that the most significant defect of
the original Batchelor argument became most apparent. So we have to accept that even
George Batchelor, the doyen of turbulence in the 1950s, could be seriously misled by a
flawed analogy. Even Homer nods!
‘isovortical perturbations’ for which ω (and not the analogous field u) is frozen in the fluid
(see § 9 below).
∂θ/∂t + u · ∇θ = κ∇ 2 θ, (7.6)
and the equation for the vector potential A = (0, 0, A(x, y, t)) of a two-dimensional
magnetic field B = ∇ × A,
∂A/∂t + u · ∇A = η∇ 2 A, (7.7)
when the velocity field u = (u(x, y, t), v(x, y, t), 0) is also two-dimensional. The analogy
is at its most evident in comparing the problem of mixing of a scalar field and the
problem of flux expulsion in magnetohydrodynamics. In the latter context, it was shown
through simple example (Moffatt & Kamkar 1983) that, when the magnetic Reynolds
914 P1-30
Some topological aspects of fluid dynamics
(a) (b) (c) (d)
4 4 4 4
2 2 2 2
0 0 0 0
–2 –2 –2 –2
–4 –4 –4 –4
0 2 4 6 8 10 12 0 2 4 6 8 10 12 0 2 4 6 8 10 12 0 2 4 6 8 10 12
Figure 16. Contours A = const. (i.e. B-lines), when the field A(x, y, t) is distorted by the shear flow
u = (tanh y, 0, 0) with initial condition A(x, y, 0) = sin x; Rm = 2000; (a) t = 10; reconnection (change of
topology) is evidenced by the field lines that do not cross the plane y = 0; (b) t = 150; the closed field loops
provide further evidence of reconnection; (c) t = 500; most of the B-lines have now reconnected and the field
in the region |y| 2 is weak; (d) t = 750; flux expulsion is now at an advanced stage; the field continues to
decay on the ohmic time scale O(Rm ) in the region |y| 2.
(a) (b) 6
1.2
5
1.0
4
0.8
0.6 3
0.4 2
0.2 1
y y
–5 0 5 –6 –4 –2 0 2 4 6
Figure 17. Profiles of (a) B2y ≡ (∂A/∂x)2 and (b) B2x ≡ (∂A/∂y)2 for t = 10 (green), 150 (blue), 500 (red) and
750 (purple).
number Rm = VL/η is sufficiently large, the relevant time scale for flux expulsion is
O(Rm 1/3 L/V), much less than the natural ohmic diffusion time scale O(Rm L/V). The
corresponding time scale for ‘homogenisation’ of a scalar field within the shearing region
is then T1 = O(Pe1/3 L/V) where Pe is the Péclet number Pe = VL/κ. Rhines & Young
(1983) showed that the situation can be much more complicated when the flow has regions
https://doi.org/10.1017/jfm.2020.230 Published online by Cambridge University Press
of closed streamlines, in that, although the initial process of homogenisation does occur on
the short time scale T1 , significant scalar field variations can survive on the much longer
time scale T2 = O(PeL/V), in regions where lines of constant θ coincide with the closed
streamlines of the flow.
The complexity of the situation is illustrated in figure 16 which shows B-lines (i.e.
contours A = const., equivalently θ = const.) for field evolution under the imposed
shearing flow u = (tanh y, 0, 0) (in dimensionless form). The initial condition is
A(x, y, 0) = sin x, so the field is initially in the y-direction, and sinusoidal in x. In
this computation, Rm = 2000. Reconnection of field lines A = const. involving obvious
change of topology is evident at early times in the shear region |y| 2 in the
neighbourhood of the ‘zero field line’ A = 1. As time advances, the complexity of the field
in the shear region increases as reconnection continues, flux being continuously expelled,
as indicated in figure 17; by the time t = 500, the field is very weak in this region, and by
t = 750 it is almost totally expelled, the field continuing to decay on the ohmic time scale
in the external region |y| 2. Since here Rm 1/3 ∼ 12, it would seem that Rm is not nearly
large enough at 2000 for the asymptotic flux expulsion time scale Rm 1/3 to be realised.
914 P1-31
H.K. Moffatt
For the corresponding problem for the scalar field θ at Péclet number Pe = 2000, the
field is effectively homogenised in the shearing region by the same time t ∼ 750.
8. Magnetic relaxation
Magnetic relaxation is a means by which magnetostatic equilibria of arbitrarily prescribed
field topology may be determined. Through the B-u analogy (§ 7.2), this is also a means of
determining steady Euler flows of arbitrary streamline topology. The idea is to construct
a process in which the magnetic energy decreases monotonically, while the field topology
is conserved. Obviously the frozen-field equation,
∂B
= ∇ × (u × B), (8.1)
∂t
must be part of this process; we can then be sure that the topology of B is conserved, at
least for so long as the field remains smooth. We need to couple this with an equation
for u, that will guarantee decrease of energy. The process will
lead to a non-trivial result
only if there is a lower bound for magnetic energy M(t) = B2 dV/2, the integral being
through the fluid domain. This lower bound is provided by an inequality first stated by
Arnold (1974).
this gives
A dV ≤ LD B2 dV,
2 2
(8.3)
where LD is the maximum span of the domain D. Combining (8.2) and (8.3) leads
immediately to the inequality
−1
2M(t) ≥ LD |HM |, (8.4)
so that, provided HM = / 0, this provides the require lower bound on M(t). Moreover
equality occurs here only if B is everywhere parallel to A, i.e. only if A is a Beltrami
field throughout D.
The physical reason for the Arnold inequality (8.4) is clear: the condition HM = / 0
implies a non-trivial topology of the B-field, in that there must exist some field lines
in D that cannot be shrunk to a point without crossing other field lines, and this is
why the magnetic energy cannot be reduced to zero by any kinematically possible
volume-preserving flow.
914 P1-32
Some topological aspects of fluid dynamics
(a) (b)
Figure 18. Links for which the link helicity is zero. (a) The Whitehead link; (b) the Borromean rings.
that it cannot be separated into disjoint subfields which could in principle be swept far apart
from each other. If the support of such a connected field is deformed by fluid flow (e.g.
by uniform irrotational strain) to an extent far greater than 2R, then, although there may
be an initial decrease of energy, this will ultimately increase due to field-line stretching.
At any rate, we can define a radius R1 , say, (≥ R), such that the field energy increases if
the support of the field is stretched to a span greater than 2R1 . Energy reduction can then
only be achieved by contraction. It is then reasonable to assert that Arnold’s inequality
still holds with LD replaced by R1 ; a formal proof of this assertion should not be difficult.
thus exhibiting the desired monotonic decrease of M(t). This decrease will persist until
u ≡ 0, and then, with BE (x) = lim B(x, t) and similarly for j and p, (8.6) gives
jE × BE = ∇pE , (8.8)
the required equation of magnetostatic equilibrium, for which M(t) is minimised.
Throughout this relaxation process, which will generally run to t = ∞, the field B(x, t)
is a frozen-in field, so that all links and knots of flux tubes are conserved. For finite t,
the flow defines an ‘isotopy’, i.e. a continuous time-dependent mapping of the field of
fluid particles via the Lagrangian displacement field a → X (a, t), where a = X (a, 0). But
here, care is needed, because in the limit t → ∞ (and only then is the equilibrium (8.8)
attained) there is no guarantee that the mapping remains continuous, and indeed tangential
field discontinuities are to be expected, as illustrated by the prototypical case of the Hopf
link (§ 8.2.2 below).
https://doi.org/10.1017/jfm.2020.230 Published online by Cambridge University Press
B01 V1
Figure 19. Relaxation of the Hopf link: (a) two untwisted but linked flux tubes; (b) contracted state when the
tubes make contact; (c) fully relaxed axisymmetric state.
topological constraint begins to ‘bite’. But relaxation does not stop here; for field lines may
still be shortened if either flux tube spreads itself round the other as in figure 19(c), there
being then two distinct asymptotic states, both axisymmetric. The discontinuity is then a
current sheet on the torus of contact.
Figure 20. Model for the formation of current sheets: movement of the end points of an initially uniform field
produces braids of arbitrary complexity (after Parker 1994).
Such chaotic Beltrami fields are not unknown; we have seen one example in the ABC
flow described in § 4.1. But here we face an extremely challenging problem: as pointed
out by Arnold (1974), there are not nearly enough Beltrami fields, analytic in x in a given
domain D, to cover every conceivable topology that may be adopted for the initial field
B0 (x) of the relaxation process. Of course, as we have seen, discontinuities may develop;
but for chaotic fields, it is likely that such discontinuities may become densely stacked
throughout the fluid. Bajer (2005) predicts a ‘devil’s staircase’ of such discontinuities,
an intriguing idea that calls for further investigation. The problem is as challenging
numerically as it is analytically.
Figure 21. Simplified model of the Parker mechanism: two flux tubes, blue and red, span the space between
parallel planes; the planes and the fluid are assumed perfectly conducting. Relaxation causes shortening of the
red tube, and deformation of the blue tube, the end points being fixed, so that contact between the tubes is
inevitable; a current sheet forms on the area of contact.
turn requires viscosity. But even if the distribution of twist is non-uniform, relaxation will
still always lead to contact, provided the angle of twist at the upper boundary is greater
than π. In practice, winding and relaxation will occur simultaneously; if the relaxation
occurs on a much shorter time scale than the winding, then the field will be permanently
in a nearly minimum-energy state (with or without tangential discontinuities), as in effect
assumed by Parker.
Current sheets are the source of strong plasma heating due to ohmic dissipation, a
favoured mechanism for the heating of the solar corona to temperatures of order 106 K;
here, the magnetic field is ‘anchored’ at the photospheric surface, and the footpoints move
in response to sub-surface turbulence. Current sheets in the relaxed coronal field are an
inevitable consequence. Kinematic viscosity ν = μ/ρ is many orders of magnitude greater
than η in the corona because of the extremely high temperature and the low particle number
density n = 1015 m−3 ; Craig & Litvinenko (2009) estimate ν/η ∼ 1010 , although they also
warn that the viscous stress tensor is strongly anisotropic in this very diffuse magnetised
https://doi.org/10.1017/jfm.2020.230 Published online by Cambridge University Press
medium. Much work on braided magnetic fields has been stimulated by the Parker model
(see particularly Wilmot-Smith, Pontin & Hornig 2010; Pontin et al. 2011).
Figure 22. Conceptual representation of the foliations of the (infinite-dimensional) function space of the B- or
U-fields; pink represents the isomagnetic foliation; blue the isovortical foliation. Stability criteria are different
for isomagnetic or isovortical perturbations.
inequality
2 (u
1 2
+ B ) dV ≥
2
u · B dV = |HC (D)|. (8.11)
D D
A relaxation process has been constructed by Vladimirov, Moffatt & Ilin (1999)
satisfying these topological constraints. The end result of this relaxation process is a
velocity field U(x) and magnetic field B(x) that satisfy the steady equations of ideal
magnetohydrodynamics, which may be expressed in the form
Ω = βU = αJ = αβ B, (8.13)
where α and β are constants. Again, these relaxed chaotic Beltrami fields must presumably
incorporate an incalculable profusion of discontinuities.
https://doi.org/10.1017/jfm.2020.230 Published online by Cambridge University Press
9. Stability
As already indicated, the analogy (7.5a–c) is limited to steady states, but does not
extend to unsteady perturbations about these states. For the magnetostatic equilibria, such
perturbations involve frozen-field deformation of the B-field, or so-called ‘isomagnetic’
perturbations. We must imagine an ‘isomagnetic foliation’ of the function space of
B-fields, as in figure 22, a B-field being constrained to evolve on a single folium.
For the analogous Euler flow, it is the vorticity field that is frozen under perturbations
governed by the Euler equations; so we must similarly imagine an ‘isovortical foliation’ of
the same functions space, which is distinct from the isomagnetic foliation. Perturbations
on the two foliations can behave quite differently.
v(x)
− 21 ξ · ∇ξ + O(ξ3)
η
ξ(x,τ)
x
0
Figure 23. Sketch indicating the relation between the Lagrangian displacement ξ (x, τ ) induced by a virtual
velocity field v(x) and the associated Eulerian displacement η after a short time τ ; these displacement functions
are related as in (9.1).
evident from the sketch, the associated Eulerian displacement η(x), satisfying ∇ · η = 0,
is given by
η(x) = ξ − 12 ξ · ∇ξ + O(ξ )2 . (9.1)
With this displacement, the field BE (x) is distorted to B(x) = BE (x) + δ 1 B + δ 2 B +
O(|η|3 ), where
δ 1 B = ∇ × (η × BE ), δ 2 B = 12 ∇ × (η × δ 1 B), (9.2a,b)
It is not difficult to show that, by virtue of (7.3a,b), the first variation δ 1 M is zero, as might
be expected for an equilibrium state. The expression for δ 2 M may then be manipulated to
https://doi.org/10.1017/jfm.2020.230 Published online by Cambridge University Press
the form
δ M = 2 [(∇ × (η × BE ))2 − (η × jE · (∇ × (η × BE ))] dV.
2 1
(9.4)
Stability is then assured if δ 2 M ≥ for all ‘admissible displacements’ η(x), i.e. those
satisfying ∇ · η = 0 in the relevant domain D, and n · η = 0 on ∂ D.
For this result to be useful, it is necessary to put a bound on the term BE · δ 2 B dV in
(9.3a,b). This involves identifying the lowest positive eigenvalue λ of a rather complicated
eigenvalue problem, and leads to the result
2
−1/2
δ M ≥ 2 1−λ
2 1
δ B
1
dV, (9.5)
so that the condition δ 2 M ≥ 0 simply requires that λ ≥ 1. A result of this kind was first
found by Bernstein et al. (1958); details of the above derivation may be found in Moffatt
(1986). Of course, as already observed, magnetostatic equilibria obtained by the method of
magnetic relaxation are stable, and must then automatically satisfy the condition δ 2 M ≥ 0.
914 P1-39
H.K. Moffatt
9.2. Instability of analogous Euler flows
Consider now a steady Euler flow uE (x) in D, with kinetic energy KE = u2E dV/2. Just
as for the magnetostatic equilibrium, we may (following Arnold 1965b) investigate the
stability of this flow by considering the second variation of uE (x), but now with respect
to the isovortical perturbations permitted by the Euler equations. It is now therefore the
vorticity ω = ∇ × u that is given by ω(x) = ωE (x) + δ 1 ω + δ 2 ω + O(|η|3 ), where
δ 1 ω = ∇ × (η × ωE ), δ 2 ω = 12 ∇ × (η × δ 1 ω), (9.6a,b)
δ 1 u = (η × ωE )s , δ 2 u = 12 (η × δ 1 ω)s , (9.7a,b)
and where the suffix s denotes the ‘solenoidal projection’, required to ensure that ∇ ·
δ 1 u = ∇ · δ 2 u = 0.
Again it may be shown that the first variation δ 1 K of kinetic energy is zero by virtue of
the equilibrium condition (7.4a–c); and for admissible η the second variation δ 2 K may be
expressed in the form
δ K = 2 [(δ 1 u)2 + 2uE · δ 2 u] dV
2 1
= 12 [(η × ωE )2s − (η × ωE )s · ∇ × (η × uE )] dV. (9.8)
There is an obvious difference between this expression for δ 2 K and δ 2 M as given by (9.4);
it is this difference that accounts for the different stability properties of magnetostatic
equilibria and their analogous Euler flows. This is most easily illustrated by the following
simple example.
in cylindrical polar coordinates (r, θ, z). We may take D to be the domain a1 <
r < a2 , 0 < z < z1 , the boundary ∂ D being perfectly conducting. We may consider
the stability of these states to admissible axisymmetric disturbances of the form
η = (ηr , 0, ηz ) = r−1 (∂ψ/∂z, 0, −∂ψ/∂r), with ψ = 0 on ∂ D. The expressions (9.4) for
δ 2 M and (9.8) for δ 2 K may in this case be reduced and manipulated to the form
2 d b 2 2 d
δ M = −π
2
ηr r dr dz, δ K = 2π
2
ηr2 (rv)2 r−2 dr dz. (9.10a,b)
dr r dr
Thus, if (b/r)2 is a monotonic decreasing function of r throughout the range a1 < r < a2 ,
then M is minimal at M = ME , and so the magnetostatic equilibrium BE is stable. We
may note that this result may be most simply obtained by considering the energy needed
to interchange two flux tubes of the same volume at different radii r1 < r2 , say, the flux
in each tube being constant: the energy needed for the interchange is then proportional
to [b(r1 )/r1 ]2 − [b(r2 )/r2 ]2 , and if this is positive, the equilibrium is stable, at least to
axisymmetric perturbations of the kind considered.
914 P1-40
Some topological aspects of fluid dynamics
Similarly, K is minimal at K = KE if (rv)2 is an increasing function of r, i.e. if
the circulation 2π|rv(r)| increases outwards; this is just the sufficient condition for
stability known as Rayleigh’s criterion (Rayleigh 1917). Equally, a necessary condition for
instability of the analogous Euler flow is that the circulation should decrease outwards.
Thus for example, if both b(r) and v(r) are proportional to r−3/2 , then (b(r)/r)2
and (rv(r))2 are both monotonic decreasing functions of r, so that the magnetostatic
equilibrium is stable, whereas the analogous Euler flow satisfies the necessary condition
for instability.
Examiner for the Mathematical Tripos at Cambridge University in 1866. This may be
found in Volume II of the Scientific Papers and Letters of James Clerk Maxwell (Harman
1995) as a ‘Draft question on the stability of vortex motion’:
A mass M of fluid is running round a circular groove or channel of radius a with velocity u. An
equal mass is running round another channel of radius b with velocity v. The one channel is made
to expand and the other to contract till their radii are exchanged. Show that the work expended in
effecting the change is
1 u2 v2
− − 2 (a2 − b2 )M.
2 b2 a
Hence show that the motion of a fluid in a circular whirlpool will be stable or unstable according as
the areas described by particles in equal times increase or diminish from centre to circumference.
The final sentence of this question (the sting in the tail!) is particularly noteworthy, as
being essentially equivalent to Rayleigh’s criterion, evidently known to Maxwell some
50 years earlier! The word ‘circulation’ had not entered the literature of hydrodynamics
in 1866, but Maxwell’s ‘area described by a particle’ in time δt is just (r2 (v/r)δt)/2 =
914 P1-41
H.K. Moffatt
rvδt/2, so proportional to circulation. Interesting that Maxwell should start here from a
simple thought experiment, and then broaden to a general principle!
∂u 1
+ u · ∇U + U · ∇u = − ∇p + ν∇ 2 u,
https://doi.org/10.1017/jfm.2020.230 Published online by Cambridge University Press
(9.11)
∂t ρ
where p(x, t) is the pressure field and ρ is the (constant) fluid density, admits ‘Kelvin wave’
solutions of the form [u, p/ρ] = [A(t), P(t)] exp [ik(t) · x], where k(t) = (k01 , k02 −
αt k01 , k03 ). The effect of the shear is to tilt the wave fronts progressively in a clockwise
sense until the normal to these wavefronts is nearly parallel to the y-axis.
The effect of viscosity is to provide a damping factor D(t) for the wave amplitude, where
D(t) = exp −ν k (t) dt = exp −ν k02 t − k01 k02 αt2 + 13 k01
2
α t
2 2 3
. (9.12)
This damping is ultimately responsible for the decay of the wave disturbance; if k01 = / 0,
then for large αt, D(t) ∼ exp[−(νk01 2 α 2 t3 /3)], so that the wave decays on a time scale
2 α 2 )−1/3 (cf. the Rhines & Young time scale encountered in § 7.2). We assume
tν = (νk01
here that ν is small, and focus on the development for t tν ; during this stage, we may
simply set ν = 0.
914 P1-42
Some topological aspects of fluid dynamics
(a) (b)
αt
20 40 60 80 100
αt –5
10 20 30 40 50
–10
–10 –15
–20
–20
–25
–30 –30
Figure 24. Transient instability; k20 = 1, k30 = 0.5; A1 (0) = 0.11, A2 (0) = 0.5 (a) ε = 0; k10 = 0.3, (red),
0.1 (blue); A1 (t) (solid), A2 (t) (dashed), A3 (t) (dotted); the transient instability shows in A1 (t) and becomes
more marked as k10 decreases. (b) Effect of viscosity; k10 = 0.1; ε = Re−1 = 10−2 (red), 10−3 (green), 10−4
(purple), 0 (blue); as ε decreases, the curves shadow the inviscid limit (blue) for an increasing period of time,
but ultimately diverge from it and asymptote to zero.
Elimination of the pressure term leads to a dynamical system for the amplitude
components
dA1 2
2αA2 k01 dA2 2αA2 k01 (k02 − αtk01 ) dA3 2αA2 k01 k03
+ αA2 = , = , = ,
dt |k(t)|2 dt |k(t)|2 dt |k(t)|2
(9.13a–c)
which may be solved explicitly for the components A1 (t), A2 (t), A3 (t). Typical behaviour is
shown in figure 24(a) for initial conditions k(0) = (0.3, 1, 0.5), A(0) = (0.1, 0.5, −0.53)
(red) and k(0) = (0.1, 1, 0.5), A(0) = (0.1, 0.5, −0.51) (blue). It is evident that A1 (t)
exhibits a relatively long period of linear growth when k10 is small; this is the key
characteristic of transient growth.
Figure 24(b) shows the effect of viscosity on the component A1 (t) for the case k10 = 0.1
through the dimensionless parameter ε = νk02 2 /α (an inverse Reynolds number) when the
decay factor D(t) is restored. As ε decreases, the solution shadows the inviscid solution
(shown again in blue) more and more closely, but ultimately diverges from this solution
and decays to zero through the effect of viscosity.
If the perturbation vorticity is random it can be Fourier analysed, and it is then the
ingredients with k1 small that are subject to this type of transient growth, the associated
https://doi.org/10.1017/jfm.2020.230 Published online by Cambridge University Press
that A ∼ V/L must increase as L decreases. Hence, if the knot K is non-trivial, it must
eventually come into contact with itself. It is at this stage that the knot topology ‘bites’
leading ultimately to a magnetostatic equilibrium in which L ∼ V 1/3 and A ∼ V 2/3 . In this
equilibrium, the magnetic energy is minimal and proportional to Φ 2 V −1/3 . This minimum
energy Mmin can (and does) depend also on h; hence we must have
(a) (b)
Flux Volume V
L ~ V1/3
A ~ V2/3
Figure 25. (a) Schematic of the relaxation of the trefoil knot T2,3 to its minimum energy state; and (b) of the
same knot in the configuration T3,2 to a different minimum-energy state.
160
140
T2,9
120
m (h)
100 T2,7
80
T2,5
60
T2,3
40
0 5 10 15 20
h
Figure 26. Minimum-energy curves for four torus knots; the minima occur for h > 0, reflecting knot
chirality; for small h, m(h) increases with increasing knot complexity. (From Chui & Moffatt 1995.)
There are, however, two important differences here in that (i) the tube cross-section
is assumed circular, whereas in magnetic relaxation, the cross-section can deform
particularly when the tube comes into contact with itself; and (ii) there is no counterpart
of twist and associated helicity in the simple knot-tightening process, which perhaps best
resembles the magnetic tube with the particular value of h for which m(h) is minimised.
Computation of LR is somewhat easier for these reasons, and great progress has been
made in determining LR and the corresponding tight-knot configurations for more than
350 distinct knots and links by the method of ‘constrained gradient descent’ (Ashton et al.
2011).
Figure 27 from Cantarella et al. (2014) shows three tight configurations of the torus knot
T2,5 ; the first has fivefold symmetry; the second (in the alternative configuration T5,2 )
has twofold symmetry, and considerably greater normalised rope length. The third is a
perturbation of the first, breaking the fivefold symmetry in a way that allows the knot to
occupy a smaller volume, thus slightly reducing LR . This is entirely consistent with the
presumed existence of an ‘energy spectrum’, if energy is here identified with normalised
rope length. It is perhaps no more than coincidental that the value LR = 48.23 is quite
close to the value mmin ≈ 48 for the magnetic torus knot T2,5 , as shown by figure 26.
914 P1-46
Some topological aspects of fluid dynamics
(a) (b) (c)
Figure 27. Three tight configurations of the torus knot T2,5 ; the normalised rope lengths are as indicated, and
it is notable that the least of these is for the configuration on the right that breaks both symmetries. (a) Fivefold
symmetry, LR = 48.23; (b) twofold symmetry, LR = 62.56; (c) both symmetries broken, LR = 47.21. (From
Cantarella et al. 2014, with permission.)
Figure 28. Schematic diagram of the trefoil vortex just before reconnection; arrows indicate the direction of
the vorticity; plan view on the left, side view on the right, showing the three stretched, antiparallel, skewed
vortex pairs. The upper ring propagates upwards more rapidly than the lower, providing the ‘forced’ stretching
of the vortex pairs.
the desired knotted vortex. Such knotted airfoils were created by 3D printer. In principle
any knot or link can be created by this technique. Although the possible existence of
knotted vortices had been hypothesised since the time of Kelvin (Thomson 1869) and
although linkage of vortex rings had been computationally achieved in recent decades
(e.g. Aref & Zawadzki 1991), this was the first time that such vortices had been realised
experimentally, thus giving tremendous impetus to research in this area.
Figure 28 shows a schematic diagram of the trefoil vortex in the configuration created
by Kleckner & Irvine (2013). The inner loop of the vortex propagates more rapidly that
the outer loop, leading in the experiment to the stretching of three anti-parallel vortex
pairs. The vortex is, as anticipated, highly unstable, the threefold symmetry being broken,
so that one of these vortex pairs extends more rapidly than the other two, and in a quite
irregular manner. Just as for the stretched Burgers vortex, the stretching here involves a
compensating inflow which decreases the minimum separation 2s of the two oppositely
directed strands. When s is decreased to the vortex cross-sectional scale δ, reconnection
takes place in an apparently explosive manner, effectively eliminating this ‘strand pair’,
and changing the topology of the trefoil to that of the unknot. A subsequent reconnection
of one of the other strand pairs changes the topology again to that of two disjoint unknots.
914 P1-47
H.K. Moffatt
(b) 1.0 y
0.5
(a) 2s0 0–0.5
–1.0
0
–Γ Γ
δ0 –0.5
R z
–1.0
C1 C2
x=0
–1 0 1
x
Figure 29. (a) Two counter-rotating circular vortices of radius R = κ0−1 propagate towards each other at angle
2α; (b) early deformation of the upper part of the vortices, the dashed curves indicating the positions at t = 0.
(After Moffatt & Kimura (2019a).)
This reconnection process was analysed by Kimura & Moffatt (2014), taking account
only of the strain field associated with the stretching, but neglecting the mutual interaction
of the two linear vortex filaments (a neglect similar to that adopted in ‘rapid distortion
theory’). The resulting linear treatment captured some, though by no means all, of the
nature of the vortex reconnection process. In particular, it was found that the helicity
associated with the initial writhe of the skewed vortices is destroyed and is not replaced by
twist; thus, helicity can apparently be destroyed on the reconnection time scale.
In a subsequent experiment, Scheeler et al. (2017) have succeeded in measuring the
helicity in a helical vortex produced in water by the above technique. The vortex was
visualised by coloured dye released from a sharp helical trailing edge, with additional
dye blobs distributed periodically along the axis of the helix. The blobs are shed into the
vortex and are trapped in its core. Measurement of the velocity of these blobs gives a direct
measure of the component of velocity parallel to the vorticity, and hence of the helicity of
the flow. By this technique, it was found that the helicity does decay (although there is here
no question of reconnection), whereas the writhe of the vortex remains nearly constant on
a much longer time scale. The inference is that viscosity dissipates the twist, and only the
writhe contribution to helicity survives. Whether this is a general tendency for a random
vorticity distribution, as in turbulent flow, is as yet unclear.
https://doi.org/10.1017/jfm.2020.230 Published online by Cambridge University Press
–γ(τ)Γ γ(τ)Γ
Figure 30. Pyramid reconnection: the incident (blue) vortices reconnect in the immediate neighbourhood of
the vertex of the pyramid, an increasing proportion of their circulations emerging as the downward-propagating
(red) ‘ejected’ vortices. Note the compatibility of the vorticity directions, indicated by the arrows in the incident
and ejected vortices.
When reconnection starts, the situation very near the vertex of the pyramid is as sketched
in figure 30. Here, a proportion of the incident (blue) vorticity has been reconnected,
giving rise to the downward-propagating (red) vortices (a ‘bridging’ process, since the red
vortices bridge the gap between the incident vortices). During reconnection, the incident
vortices retain a proportion γ (t) of their initial circulation, where γ (t) must decrease from
1 to very near zero. We return to this ‘pyramid reconnection’ process in § 11.3 below.
Navier–Stokes equations. In a statistically steady state, one must regard as the rate
of supply of energy to the turbulence on the ‘energy-containing scale’ 0 , as well as
its subsequent cascade to smaller scales. It is then a given parameter of the turbulence,
independent of ν. The (Kolmogorov) scale at which energy is dissipated depends only
on and ν, and is determined on dimensional grounds as ν ∼ (ν 3 /)1/4 0 . Thus, as
ν → 0, ν → 0 also, and the turbulence develops ever finer structure in this limit.
This at any rate is the traditional picture. However, an inconsistency in the theory was
recognised by Kolmogorov (1962) himself, in that the unaveraged quantity Ω(x, t) = ω2
is by no means uniform throughout the flow field; and in regions where Ω is greater than
its average, the energy cascade (and the associated vortex stretching) will progress more
vigorously, a process that can rapidly lead to intermittency, i.e. extreme ‘spottiness’ in the
distribution of Ω.
This raises a question of central and acute importance in the theory of turbulence: given
that in (11.1) remains prescribed and constant in the limit ν → 0, what is the nature of
the typical singularity of ω2 in this limit, required to satisfy this equation? This question
is closely related to the fundamental question of regularity of the Navier–Stokes equation
first raised by Leray (1934), a question that can be most simply phrased as follows: given
914 P1-49
H.K. Moffatt
at some initial instant a smooth localised velocity field u0 (x) of finite kinetic energy in an
incompressible fluid of arbitrarily small ν filling all space, can a singularity of the field
appear within a finite time under evolution governed by the Navier–Stokes equations?
The widely held belief is that the Navier–Stokes equations are regular, in the sense that
such finite-time singularities cannot occur. But is this not then in conflict with (11.1),
which requires some kind of singularity at least in the limit ν → 0? We now address this
troublesome issue.
inappropriate to reproduce the details here. Enough to say that, under certain admittedly
optimistic simplifications (neglect of core deformation of the incident blue vortices in
figure 30 during the reconnection process, and neglect of the effect of the receding red
vortices on the continuing progress of the incident blue vortices towards the pyramid apex),
the behaviour is found to be described by a nonlinear dynamical system involving four
functions of the dimensionless time τ = (Γ /R2 )t, namely the separation parameter s(τ ),
the vortex curvature κ(τ ) and scale of cross-section δ(τ ) at the points of nearest approach
(the ‘tipping points’), and the fraction γ (τ ) of unreconnected (blue) vorticity at time τ . In
dimensionless form, this dynamical system is
ds κ cos α s dκ κ cos α sin α d δ2 κ cos α 2
= −γ log + β1 , =γ , =ε−γ δ ,
dτ 4π δ dτ 4πs 2 dτ 4πs
(11.2a–c)
and
dγ sγ
= −ε √ 3 exp [−s2 /4δ 2 ], (11.3)
dτ 2 πδ
where α = π/4, β1 = 0.4417 and ε = ν/Γ 1, an inverse vortex Reynolds number.
914 P1-50
Some topological aspects of fluid dynamics
(a) (b)
1.2 γ(τ) 100
1.0
80
0.8
δ(τ)/δ(0) 60
0.6
40
0.4
ω(τ)/ω(0)
0.2 20
τ τ
0 0.05 0.10 0.15 0.20 0.25 0.240 0.241 0.242 0.243 0.244 0.245
Figure 31. Evolution determined by (11.2a–c) and (11.3) with initial conditions s(0) = 0.1, δ(0) = 0.01,
κ(0) = γ (0) = 1; ε = 10−5 ; computed by Mathematica with 56-point precision; (a) the vortex-core scale
δ which appears to fall to zero at a critical time τ = τc ≈ 0.243777 . . ., at which stage γ ≈ 0.9992 . . .;
(b) corresponding late-stage evolution of ω(τ ) = γ (τ )/δ(τ )2 which appears to go to infinity at τ ≈ τc .
The interpretation of these equations is as follows. The first represents the rate of
decrease of s due to the self-induced velocity of each vortex together with an additional
contribution of opposite sign due to mutual repulsion of the vortices. The second equation
represents the rate of increase of tip curvature κ, largely due to the decrease in s and
the resulting accelerating upward movement of the tipping points. The third equation
represent the evolution of δ 2 , partly an increase due to simple viscous diffusion, and partly
a decrease due to stretching which dominates when γ ≈ 1 (i.e. before reconnection) and
κδ 2 /s ε. The fourth equation (11.3) represents the decrease in γ , which kicks in only
when s/δ has decreased to O(1), i.e. when the viscous reconnection process gets underway.
It is important to note that the small parameter ε appears in two places in this dynamical
system, which makes asymptotic analysis for the limit ε → 0 particularly challenging.
Direct numerical simulation of the reconnection process starting from the same initial
conditions as represented in figure 29(a) (Yao & Hussain 2020) has shown that (11.3)
seriously underestimates the rate of transfer of circulation from the incident to the
reconnected vortices. Nevertheless, numerical computation of the evolution governed by
(11.2a–c) and (11.3) from appropriate initial conditions is in principle straightforward, and
https://doi.org/10.1017/jfm.2020.230 Published online by Cambridge University Press
is illuminating as regards possible behaviour at very high Reynolds numbers beyond the
reach of DNS. Figure 31 shows the sort of results that can be obtained by such integration.
Here, I have chosen the vortex Reynolds number RΓ = Γ /ν = 105 , so = 10−5 . The
initial conditions were as shown in the caption. The Mathematica programme used here
stalls at an apparent singularity at τ = τc ≈ 0.243777 . . ., where δ(τ ) appears to go to
zero and the vorticity function ω(τ ) = γ (τ )/δ(τ )2 appears to go to infinity. It is tempting
to claim a finite-time singularity at τ = τc , and it is indeed very nearly a singularity. But
numerical resolution of the behaviour in the immediate vicinity of τ = τc is inadequate to
resolve this apparent singularity, whose very existence must for the time being remain a
matter of speculation.
What, it may be asked, has this to do with topology? Well, it is only the reconnection
process that can and does change vortex topology. Thus for example, reconnection of the
two circular vortices of figure 29 replaces these by a single contorted loop. In turbulent
flow, it is not unreasonable to suppose that this type of vortex reconnection is a frequent
occurrence. Each such ‘reconnection event’ is extremely localised and of extremely short
duration. The rate of dissipation of energy during the event is extremely high, but the total
energy dissipated is just the decrease in the local energy resulting from the reconnection,
914 P1-51
H.K. Moffatt
and is obviously finite. The reconnection events are random in both space and time, and the
mean number per unit volume per unit time must be just sufficient to dissipate the energy
at the rate at which it cascades from larger scales. This provides a possible scenario for
an approach to the problem of intermittency (see § 11.2 above). An approach developing
similar ideas has been recently proposed by Pomeau, Le Berre & Lehner (2019).
(i) For the corner flow problem treated in § 3.2, how does the behaviour depend on the
Reynolds number based on the remote driving mechanism? In particular how do the
positions of the first points of separation and reattachment depend on this Reynolds
number; and what then is the Reynolds number based on the scale and maximum
velocity in the second eddy of the sequence? Are we already at this stage into the
‘low-Reynolds-number’ regime?
(ii) For the free-surface-singularity problem of § 3.4, again what is the modification
in behaviour as the Reynolds number based on flow velocity and scale at the free
surface increases? For the problem of a steady laminar stream of water from a tap
into a deep bath of water, what is the critical Reynolds number and/or capillary
number at which bubbles are first entrained through the circular cusp around the
point of impact (with evident change of free-surface topology)?
(iii) For the flows of § 4 exhibiting ‘Lagrangian chaos’, what is the dependence of the
Lyapunov exponent (a measure of mixing efficiency) on the parameters determining
https://doi.org/10.1017/jfm.2020.230 Published online by Cambridge University Press
these flows? Are the flows (4.4a–c) capable of dynamo action, if the fluid is
conducting?
(iv) For the scalar field problem discussed in § 5.1, how does the maximum attained by
G2 depend on the Péclet number?
(v) For the corresponding magnetic field problem (i.e. the small-scale dynamo problem
in reflectionally symmetric turbulence), if the magnetic energy grows (or decays)
exponentially how exactly does the growth rate depend on the magnetic Prandtl
number Pm = ν/η?
(vi) For the flux expulsion problem (§ 7.3), how large does Rm really have to be before
the asymptotic high Rm behaviour is attained? And similarly of course for the
analogous vorticity homogenisation problem.
(vii) For the magnetic relaxation problem of § 8, when the initial field is chaotic, what
is the asymptotic structure of the relaxed field? Equivalently, what is the function
space within which this relaxed field must reside?
(viii) And finally, there is of course the ever-challenging finite-time singularity problem.
What is the true nature of the (near)-singularities that must appear in turbulence in
order to dissipate energy at a given mean rate in the limit ν → 0?
914 P1-52
Some topological aspects of fluid dynamics
Declaration of interests. The author reports no conflict of interest.
Author ORCIDs.
H.K. Moffatt http://orcid.org/0000-0003-2575-5111.
R EFERENCES
A LFVÉN , H. 1946 A new type of wave motion and its importance in solar physics. Acta Radiol. 27 (3–4),
228–242.
A REF, H. 1984 Stirring by chaotic advection. J. Fluid Mech. 143, 1–21.
A REF, H. & Z AWADZKI , I. 1991 Linking of vortex rings. Nature 354 (6348), 50–53.
A RNOLD , V.I. 1965a Sur la topologie des écoulements stationnaires des fluides parfaits. C. R. Acad. Sci. Paris
261, 17–20.
A RNOLD , V.I. 1965b Variational principle for three-dimensional steady-state flows of an ideal fluid. J. Appl.
Maths Mech. 29, 1002–1008.
A RNOLD , V.I. 1974 The asymptotic Hopf invariant and its applications. In Proceedings of the Summer School
in Differential Equations at Dilizhan, Erevan, Armenia [in Russian], pp. 229–256 (translation in Sel. Math.
Sov. 5, 327–345). Armenian Academy of Sciences.
A RNOLD , V.I. & K HESIN , B.A. 1998 Topological Methods in Hydrodynamics. Springer.
A SHTON , T., C ANTARELLA , J., P IATEK , M. & R AWDON , E.J. 2011 Knot tightening by constrained gradient
descent. Expl. Maths 20 (1), 57–90.
BAJER , K. 2005 Abundant singularities. Fluid Dyn. Res. 36, 301–317.
BAJER , K. & M OFFATT, H.K. 1990 On a class of steady confined Stokes flows with chaotic streamlines.
J. Fluid Mech. 212, 337–363.
BATCHELOR , G.K. 1950 On the spontaneous magnetic field in a conducting liquid in turbulent motion. Proc.
R. Soc. Lond. A 201, 405–416.
BATCHELOR , G.K. 1952 The effect of homogeneous turbulence on material lines and surfaces. Proc. R. Soc.
Lond. A 213, 349–366.
BATCHELOR , G.K. 1953 The Theory of Homogeneous Turbulence. Cambridge University Press.
B ERNSTEIN , I.B., F RIEMAN , E.A., K RUSKAL , M.D. & K ULSRUD , R.M. 1958 An energy principle for
hydromagnetic stability problems. Proc. R. Soc. Lond. A 244, 17–40.
B ETCHOV, R. 1961 Semi-isotropic turbulence and helicoidal flows. Phys. Fluids 4, 926–926.
BOGOYAVLENSKIJ , O. 2017 Counterexamples to Moffatt’s statements on vortex knots. Phys. Rev. E
95 (4), 043104.
B RAGINSKY, S.I. 1991 Towards a realistic theory of the geodynamo. Geophys. Astrophys. Fluid Dyn.
60 (1), 89–134.
C ANTARELLA , J., F U, J.H.G., M ASTIN , M. & ROYAL , J.E. 2014 Symmetric criticality for tight knots. J. Knot
Theory Ramifications 23 (02), 1450008.
C HUI , A.Y.K. & M OFFATT, H.K. 1995 The energy and helicity of knotted magnetic flux tubes. Proc. R. Soc.
https://doi.org/10.1017/jfm.2020.230 Published online by Cambridge University Press
M OFFATT, H.K. 1981 Some developments in the theory of turbulence. J. Fluid Mech. 106, 27–47.
M OFFATT, H.K. 1985 Magnetostatic equilibria and analogous Euler flows of arbitrarily complex topology.
Part 1. Fundamentals. J. Fluid Mech. 159, 359–378.
M OFFATT, H.K. 1986 Magnetostatic equilibria and analogous Euler flows of arbitrarily complex topology.
Part 2. Stability considerations. J. Fluid Mech. 166, 359–378.
M OFFATT, H.K. 1987 On the existence of Euler flows that are topologically accessible from a given flow. Rev.
Brasil. Ciências Mec. IX (2), 93–101.
M OFFATT, H.K. 1990 The energy spectrum of knots and links. Nature 347, 367–369.
M OFFATT, H.K. 1994 Book review of ‘Saffman, P.G. 1993 Vortex Dynamics. Cambridge University Press’.
Bull. Lond. Math. Soc. 26 (6), 621–622.
M OFFATT, H.K. 2008 Magnetostrophic turbulence and the geodynamo. In Computational Physics and New
Perspectives in Turbulence (ed. Y. Kaneda), pp. 339–346. Springer.
M OFFATT, H.K. 2017 The early years of the Journal of Fluid Mechanics. Style and international impact.
C. R. Méc. 345 (7), 498–504.
M OFFATT, H.K. & D ORMY, E. 2019 Self-Exciting Fluid Dynamos. Cambridge University Press.
M OFFATT, H.K. & K AMKAR , H. 1983 On the time-scale associated with flux expulsion. In Stellar and
Planetary Magnetism (ed. A. M. Soward), pp. 91–98. Gordon and Breach.
M OFFATT, H.K. & K IMURA , Y. 2019a Towards a finite-time singularity of the Navier–Stokes equations.
Part 1. Derivation and analysis of dynamical system. J. Fluid Mech. 861, 930–967.
914 P1-54
Some topological aspects of fluid dynamics
M OFFATT, H.K. & K IMURA , Y. 2019b Towards a finite-time singularity of the Navier–Stokes equations.
Part 2. Vortex reconnection and singularity evasion. J. Fluid Mech. 870, R1. (See also Corrigendum
J. Fluid Mech., 887, E2.
M OFFATT, H.K. & L OPER , D.E. 1994 The magnetostrophic rise of a buoyant parcel in the Earth’s core.
Geophys. J. Intl 117, 394–402.
M OFFATT, H.K. & P ROCTOR , M.R.E. 1985 Topological constraints associated with fast dynamo action.
J. Fluid Mech. 154, 493–507.
M OFFATT, H.K. & R ICCA , R.L. 1992 Helicity and the Călugăreanu invariant. Proc. R. Soc. Lond. A
439, 411–429.
M OFFATT, H.K. & T SINOBER , A., ed. 1990 Topological Fluid Mechanics. Cambridge University Press.
M OREAU, J.-J. 1961 Constantes d’un îlot tourbillonnaire en fluid parfait barotrope. C. R. Hebd. Seances Acad.
Sci., Paris 252, 2810–2812.
O RSZAG , S.A. 1977 Lectures on the statistical theory of turbulence. In Fluid Dynamics, Les Houches, Juillet
1973 (ed. R. Balian & J. L. Peube), pp. 235–347. Gordon and Breach.
PARKER , E.N. 1955 Hydromagnetic dynamo models. Astrophys. J. 122, 293–314.
PARKER , E.N. 1994 Spontaneous Current Sheets in Magnetic Fields. Oxford University Press.
P OMEAU, Y., L E B ERRE , M. & L EHNER , T. 2019 A case of strong nonlinearity: intermittency in highly
turbulent flows. C. R. Méc. 347 (4), 342–356.
P ONTIN , D.I., W ILMOT-S MITH , A.L., H ORNIG , G. & GALSGAARD , K. 2011 Dynamics of braided coronal
loops II. Cascade to multiple small-scale reconnection events. Astron. Astrophys. 525, A57.
R ASSKAZOV, A., C HERTOVSKIH , R. & Z HELIGOVSKY, V. 2018 Magnetic field generation by pointwise
zero-helicity three-dimensional steady flow of an incompressible electrically conducting fluid. Phys.
Rev. E 97, 043201.
R AYLEIGH , L ORD 1917 On the dynamics of revolving fluids. Proc. R. Soc. Lond. A 93, 148–154.
R AYLEIGH , L ORD 1920 Steady motion in a corner of a viscous fluid. In Scientific Papers, vol. 6, 8–21.
R HINES , P.B. & YOUNG , W.R. 1983 How rapidly is a passive scalar mixed within closed streamlines? J. Fluid
Mech. 133, 133–145.
ROBERTS , P.H. & S TIX , M. 1971 The turbulent dynamo: a translation of a series of papers by F. Krause,
K.-H. Rädler and M. Steenbeck. In Mechanical Note 60. National Center for Atmospheric Research, CO,
USA (NCAR).
ROUCHON , P. 1991 On the Arnol’d stability criterion for steady-state flows of an ideal fluid. Eur. J. Mech.
B/Fluids 10, 651–661.
SAFFMAN , P.G. 1963 On the fine-scale structure of vector fields convected by a turbulent fluid. J. Fluid Mech.
16, 545–572.
SAFFMAN , P.G. 1993 Vortex Dynamics. Cambridge University Press.
S CHEELER , M.W., VAN R EES , W.M., K EDIA , H., K LECKNER , D. & I RVINE , W.T.M. 2017 Complete
measurement of helicity and its dynamics in vortex tubes. Science 357, 487–491.
S PITZER , L. J R . 1956 Physics of Fully Ionized Gases. Interscience.
S QUIRES , T.M. & Q UAKE , S.R. 2005 Microfluidics: fluid physics at the nanoliter scale. Rev. Mod. Phys. 77,
977–1026.
https://doi.org/10.1017/jfm.2020.230 Published online by Cambridge University Press
S T. P IERRE , M.G. 1996 On the local nature of turbulence in Earth’s outer core. Geophys. Astrophys. Fluid
Dyn. 83 (3-4), 293–306.
S TASIAK , A., K ATRITCH , V. & K AUFFMAN , L.H. 1998 Ideal Knots, vol. 19. World Scientific.
S TEENBECK , M., K RAUSE , F. & R ÄDLER , K.-H. 1966 Berechnung der mittleren Lorentz-Feldstärke für ein
elektrisch leitendes Medium in turbulenter, durch Coriolis-Kräfte beeinflusster Bewegung (in German).
Z. Naturforsch. 21a, 369–376 (translation in Roberts and Stix (1971) pp. 29–47).
S TONE , H.A., NADIM , A. & S TROGATZ , S.H. 1991 Chaotic streamlines inside drops immersed in steady
Stokes flows. J. Fluid Mech. 232, 629–646.
TAIT, P.G. 1867 On the integrals of the hydrodynamical equations, which express vortex motion. Phil. Mag.
33, 485–512 (translation of Helmholtz 1858).
TAIT, P.G. 1883 Johann Benedict Listing (obituary). Nature 27, 316–317.
TAIT, P.G. 1898 On knots, I, II, III. In Scientific Papers, vol. 1.
TANEDA , S. 1979 Visualization of separating Stokes flows. J. Phys. Soc. Japan 46, 1935–1942.
TAYLOR , G.I. 1960 Aeronautics and astronautics. In Similarity Solutions of Hydrodynamic Problems (ed.
N. Hoff & W. Vincenti), pp. 21–28. Pergamon Press.
T HOMSON , W. 1867 On vortex atoms. Phil. Mag. 34, 15–24.
T HOMSON , W. 1869 On vortex motion. Trans. R. Soc. Edinburgh 25, 217–260.
T RKAL , V. 1919 C̆asopis pro pĕstování mathematiky a fysiky (in Czech). Poznámka Hydrodyn. Vazkých
Tekutin. 48, 302–311 (translation in Czech. J. Phys. 44, 97–106).
914 P1-55
H.K. Moffatt
VAINSHTEIN , S.I. & Z EL ’ DOVICH , Y.B. 1972 Origin of magnetic fields in astrophysics. Sov. Phys. Uspekhi
15, 159–172.
V LADIMIROV, V.A., M OFFATT, H.K. & I LIN , K.I. 1999 On general transformations and variational
principles for the magnetohydrodynamics of ideal fluids. Part 4. Generalized isovorticity principle for
three-dimensional flows. J. Fluid Mech. 390, 127–150.
W ILMOT-S MITH , A.L., P ONTIN , D.I. & H ORNIG , G. 2010 Dynamics of braided coronal loops I. Onset of
magnetic reconnection. Astron. Astrophys. 516, A5.
WOLTJER , L. 1958 A theorem on force-free magnetic fields. Proc. Natl Acad. Sci. 44, 489–491.
YAO , J. & H USSAIN , F. 2020 On singularity formation via viscous vortex reconnection. J. Fluid Mech.
888, R2.
https://doi.org/10.1017/jfm.2020.230 Published online by Cambridge University Press
914 P1-56