To save content items to your account,
please confirm that you agree to abide by our usage policies.
If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account.
Find out more about saving content to .
To save content items to your Kindle, first ensure no-reply@cambridge.org
is added to your Approved Personal Document E-mail List under your Personal Document Settings
on the Manage Your Content and Devices page of your Amazon account. Then enter the ‘name’ part
of your Kindle email address below.
Find out more about saving to your Kindle.
Note you can select to save to either the @free.kindle.com or @kindle.com variations.
‘@free.kindle.com’ emails are free but can only be saved to your device when it is connected to wi-fi.
‘@kindle.com’ emails can be delivered even when you are not connected to wi-fi, but note that service fees apply.
By
P.A. Fox, High Altitude Observatory, National Center for Atmospheric Research Boulder, CO 80307-3000 USA,
M.L. Theobald, Center for Solar and Space Research, Yale University New Haven, CT 06511-6666 USA,
S. Sofia, Center for Solar and Space Research, Yale University New Haven, CT 06511-6666 USA
The detailed dynamics of the Solar dynamo presents a significant challenge to our understanding of the interaction of convection and magnetic fields in the Solar interior. In this paper we discuss certain aspects of this interaction, such as modification of convective energy transport, and turbulent dissipation of magnetic fields. The latter controls the spatial distribution of the magnetic field and its time dependence. We also discuss how these results may influence current Solar dynamo calculations.
MOTIVATION
Solar activity manifests itself in many forms but perhaps most importantly through the presence of a magnetic field. The topic of this meeting is that of dynamos, in Solar and planetary contexts. In the case of the Sun the dynamo, which seems likely to be responsible for at least part of the Solar activity we observe, acts on a global scale. That is, the period of the dynamo is 22 years (a timescale distinct from those usually encountered on the Sun), sunspots appear within latitude bands and their numbers (in terms of monthly or yearly running means) increase and decrease over one cycle. There is however, a strong asymmetry of the Solar cycle in time, i.e. the growth phase is shorter (and dependent of the amount of activity) than the decay, or descending phase. In addition, the polar field of the Sun is observed to reverse around Solar maximum, again with a distinct asymmetry between hemispheres. Despite these global-scale features, the Solar magnetic field has many spatial components (Stenflo 1991) and the majority of the magnetic flux appears in small elements.
The outer fluid core of the Earth can be considered as a fluid between two hard spheres (the internal core and the rock mantle) rotating with different but close angular velocities. In the incompressible, nonconducting almost inviscid limit a singular cylindrical surface having the radius of the internal sphere appears (the Proudman solution). A shear layer forming around this surface in the non-ideal fluid may have important implications for the geodynamo.
INTRODUCTION
The aim of this short paper is to attract attention to one feature in the Earth's fluid core. The feature is an internal shear layer induced by a relative rotation of the inner core. Large gradients of the velocity around this layer may be important for the geodynamo. Note, in particular, that in the geodynamo model-Z without an account of the inner core rotation one of the basic sources (the α-effect) is assumed to be concentrated near the core-mantle boundary (Braginsky 1993).
The inner core of the Earth can be considered as a hard iron ball of radius approximately 0.2R, where R is the Earth's radius. The rest of the planet is occupied by the outer liquid core and the rock mantle in the form of spherical shells of almost equal width, 0.4R. The other iron-rock planets (Mercury, Mars), except probably Venus, also have inner cores (Stevenson 1983). As the source of compositional convection (Loper & Roberts 1983) the inner core is apparently a necessary part of the planetary dynamo.
Turbulence plays a crucial role in dynamo processes. For example, turbulent difFusion is important for the existence of the Solar dynamo. Some turbulent phenomena may be studied with presentday measurement equipment. A number of relevant diagnostics are based on the interaction of an electromagnetic beam with plasma. Here we discuss the situation in which information on plasma properties is obtained by probing plasma with a plane polarized electromagnetic beam. It is shown that the problem of recovering statistical properties of turbulence from the line integrated data can be solved uniquely using a realistic model of plasma. Analytical expressions relating structure functions of both the random density field and random magnetic field to measured structure functions have been found. This information is of importance in studies of MHD turbulence.
STATISTICAL PROPERTIES OF PLASMA TURBULENCE
Recent measurements have shown the existence of fine-scale density structures in Tokamak plasmas (Cripwell & Costley 1991). There is also experimental evidence that the anomalous (i.e. greater than collisional) particle and energy transport may be in some circumstances due to particle drift motion caused by microturbulence. These facts make the investigation of the turbulence in Tokamaks very important. To describe the phenomena, it is useful to know statistics of random magnetic and density fields.
In this paper we discuss the statistical properties of plasma which can be studied with the so-called refractometry technique probing plasma with a plane polarized electromagnetic laser or microwave beam (Gill & Magyar 1987; Weisen et al. 1988.
In order to discuss the observations in a meaningful way, it is useful to first discuss the theory, because only then does one have a meaningful context in which to place the observations. One must note that this does not imply that the theory is well understood, because this is not necessarily the case. But this approach gives some insight as to what the key observations are, and by comparing the observations with the theory one can find the weakness in the theory which can be more thoroughly studied.
Review of the early evolution
The main sequence star transforms hydrogen into helium in the central regions of the star as its source of energy. This stage is usually referred to as hydrogen burning. As the hydrogen burns, the core gradually contracts and heats up, which produces an increase in the rate at which the hydrogen is burned. The increased burning rate closely offsets the diminished fuel supply, and the luminosity of the star does not change substantially. For the first 90% of its lifetime the star remains close to the main sequence, and may double its luminosity in this time. This is shown schematically in the H-R diagram, Fig. 1, as the motion from points A to B. The initial position on the main sequence is determined only by the mass of the star, as long as it consists mainly of hydrogen. Evolutionary tracks are shown for stars of 1.1 M⊙ and 5 M⊙. The former represent the ‘low’ mass stars and the latter represent ‘intermediate’ mass stars which will be presently more completely defined.
The Childress–Soward dynamo, which uses rotating Benard convection to maintain a magnetic field against Ohmic decay, is investigated numerically. A converged three-dimensional solution of the strong field branch is presented for very small Ekman number. For strong rotation, the system is able sustain convection and act as a dynamo even for a Rayleigh number substantially less than critical. It is found that the dominant forces tend to cancel, and that the magnitudes of the curls of the Lorentz and Coriolis forces remain virtually identical.
INTRODUCTION
Numerical computations comprise an increasingly important tool in the understanding of the Earth's dynamo, and, with the increased accessibility of supercomputers, direct, realistic simulations of the geodynamo are not far off. Any such simulation must solve the equations governing a three dimensional, rapidly rotating, dynamically consistent dynamo with Lorentz force J × B present in the dominant balance of forces. The simplest dynamo with these characteristics, first proposed by Childress & Soward (1972), uses the convective motions of rapidly rotating Benard convection to drive a dynamically consistent MHD dynamo. Computationally, the Childress–Soward dynamo has the advantage of permitting the expansion of the unknown fields in Fourier series in all directions, allowing three dimensional fast Fourier transforms (FFT's) to be used in calculating the nonlinear terms. Since no fast Legendre transform exists at the moment, the resulting programs will be faster than more realistic spherical dynamo simulations, while at the same time reflecting the important features of these models.
In this paper, the strong field branch of the Childress–Soward dynamo is investigated using direct numerical simulations.
In a report of the current status of the observational studies of star formation published in 1982, Wynn-Williams introduced the subject by boldly stating “Protostars are the Holy Grail of infrared astronomy”, one of the most (ab)used astronomical quotations ever. At the time of the review, searches had been made at IR wavelengths towards a restricted sample of objects, guided by the theoretical expectation that protostars would undergo a phase of high IR luminosity during the main accretion phase (Larson 1969). The aim of these studies, and of those that followed, was to get an unambiguous example of Spitzer's definition (1948) of a protostar as “an isolated interstellar cloud undergoing inexorable gravitational contraction to form a single star”.
This definition appears now rather restrictive, since observations have shown that the structure of star forming regions is highly complex, rarely revealing well isolated, noninteracting interstellar clouds. In addition, observations in the near infrared at high spatial resolution have shown that many young stars are not single, but do have companions (Zinnecker & Wilking 1992). However, despite great effort, resulting from improvements in instrumentation and progress in the theoretical models, the same discouraging conclusion reached by Winn-Williams in 1982 still holds true: no conclusive identification of a genuine protostar has yet been made. Nevertheless, the motivation for continuing the search using the IR band remains still unquestionable.
In these lectures I will present an overview of the main properties of the star formation process with a special emphasis on stars of low- and intermediate-mass, for which observations are the most detailed and a consistent theoretical framework has been developed and tested.
Early measurements of the interstellar extinction curve at visual wavelengths (Stebbins et al., 1939) established a broad result which has survived unchanged, namely that the amount of the extinction on a magnitude scale has an approximately linear dependence on wavenumber, i.e. Aλ ∝ 1/λ. This result was later refined (Nandy 1964a, b, 1965) to show that in a second order of approximation Aλ could be represented in a plot against 1/λ, by two straight line segments intersecting at 1/λ ≅ 2.4 (μm−1), with the segment corresponding to blue wavelengths being somewhat shallower in slope than the segment corresponding to red wavelengths. The shape of the interstellar extinction curve over the waveband 1 < λ−1 λ 3μm−1 remained more or less invariant from star to star. Grain models were thus constrained to possess a wavelength dependence of extinction that accorded with this general result.
Another observational criterion that was available from the 1930's related to the amount of the interstellar extinction per unit path length. For instance at the visual wavelength, corresponding to 1/λ = 1.8 (μm)−1, the mean extinction of starlight in directions close to the galactic plane is about 2 mag/kpc.
A further result of relevance is the so-called Oort limit for the total mass density of interstellar material – gas and dust – which amounts to ˜ 3 × 10−24 g cm−3 (Oort, 1932, 1952). The dust grain density had certainly to be less than this value, probably considerably less, in view of the fact that the bulk of cosmic material is made up of H which cannot by itself condense into solid grains at a temperature above that of the cosmic microwave background, ˜ 2.7°K.
Recent results concerning the amplification of magnetic field frozen to a two-dimensional spatially periodic flow consisting of two distinct pulsed Beltrami waves are summarised. The period a of each pulse is long (α ≫ 1) so that fluid particles make excursions large compared to the periodicity length. The action of the flow is reduced to a map T of a complex vector field Z measuring the magnetic field at the end of each pulse. Attention is focused on the mean field (Z) produced. Under the assumption, (Tk+2Z) − |λ∈|2«TkZ) → 0 as K → ∈, an asymptotic representation of the complex constant λ∈ is obtained, which determines the growth rate α−1(α|λ∈|). The main result is the construction of a family of smooth vector fields ZN and complex constants λN with the properties (for even N), and for all integers K(> 0), where ∈ = α−3/2. The relation of ZN and λN to the modes of the corresponding dissipative problem with the fastest growth rates is discussed.
INTRODUCTION
The key characteristic of a fluid motion necessary for fast dynamo action is the existence of a positive Liapunov exponent. Childress (1992) calls a motion with this property a stretching flow and it is generally manifest by chaotic particle paths. For steady flows the regions of exponential stretching are often small, as they are, for example, in the case of the spatially periodic flows discussed by Dombre et al. (1986). The numerical demonstration of fast dynamos in such flows has proved difficult and Galloway & Frisch's (1984, 1986) results were inconclusive even at the largest values of the magnetic Reynolds number reached.
By
D.J. Galloway, School of Mathematics and Statistics, University of Sydney, Sydney, NSW 2006 Australia,
N.R. O'Brian, School of Mathematics and Statistics, University of Sydney, Sydney, NSW 2006 Australia
We demonstrate the existence of a two-dimensional incompressible flow having a negative and isotropic eddy viscosity. Here, we understand by ‘eddy viscosity’ the sum of the molecular viscosity and of the small-scale flow contribution. The flow is deterministic, time-independent, space-periodic and has φ/3 rotational invariance. The eddy viscosity is calculated by multiscale techniques. The resulting equations for the transport coefficients are solved (i) by a Pade-resummed Reynolds number expansion and (ii) by direct numerical simulation. Results agree completely.
It is known that the action of a small-scale incompressible flow (having suitable symmetries) on a large-scale perturbation of small amplitude is ‘formally’ diffusive (Kraichnan 1976; Dubrulle & Frisch 1991). There are two essential assumptions. The first one is scale-separation: the ratio e between the typical length-scale of the basic flow and that of the perturbation is small. The second one is the absence of a large-scale AKA effect (Frisch et al 1987). If the basic flow is parity-invariant (i.e. has a center of symmetry), this condition is automatically satisfied. By ‘formally’ diffusive, we understand that, unlike the case of the eddy diffusivity for a passive scalar (Frisch 1989), the eddy viscosity tensor need not be positive definite. There are indeed examples of strongly anisotropic flows (e.g. the Kolmogorov flow), where some components of the tensor are negative, resulting in a large-scale instability (Meshalkin & Sinai 1961; Green 1974; Sivashinsky 1985; Sivashinsky & Yakhot 1985).
When the eddy viscosity tensor is isotropic, the equation for the perturbation reduces to an ordinary diffusion equation, with diffusion coefficient uE.
By
A. Brandenburg, Isaac Newton Institute for Mathematical Sciences, University of Cambridge, 20 Clarkson Rd., Cambridge, CB3 OEH UK,
A. Brandenburg, Isaac Newton Institute for Mathematical Sciences, University of Cambridge, 20 Clarkson Rd., Cambridge, CB3 OEH UK,
I. Procaccia,
D. Segel, Department of Chemical Physics, The Weizmann Institute of Science, Rehovot 76100, Israel,
A. Vincent,
M. Manzini, CERFACS, 42 Avenue Coriolis, F-31057 Toulouse, France
The infrared region of the electromagnetic spectrum spans a large range in wavelengths compared to that of normal visible light and it has not been easy to develop technologies which allow astronomers to study the infrared waveband. However, in the last 8 – 10 years there has been a tremendous growth in the field of infrared astronomy. This growth has been stimulated in part by the construction of infrared telescopes and by successful space missions, but the most important event has been the development of very sensitive imaging devices called infrared arrays. Before describing these detectors and their uses in instrumentation, it is useful to begin with a brief historical review of infrared astronomy and an explanation of the terminology of the subject.
Historical review: from Herschel to IRAS
Infrared astronomy had an early beginning when, in 1800, Sir William Herschel noted that a thermometer placed just beyond the red end of the optical spectrum of the Sun registered an increase in temperature due to the presence of invisible radiation which he called calorific rays. He even demonstrated that these rays were reflected and refracted like ordinary light. This discovery came 65 years before James Clerk Maxwell's theory on the existence of an entire spectrum of electromagnetic radiation.
Despite that early start, and some additional development of infrared detectors by Edison, and later by Golay, no major breakthroughs in infrared astronomy occurred until the 1950s — the era of the transistor — when simple, photoelectric detectors made from semiconductor crystals became possible.
We discuss the consequences of nonlinear effects on the effective magnetic field transport coefficients in a magnetofluid; such transport effects lie at the heart of modern astrophysical dynamo theories. The particular focus of our discussion is on the distinction between fully turbulent and quasi-steady flows; we show that these two types of flows both show suppression of effective magnetic field transport, but are distinguished by the amplitude of the suppression effect: suppression is substantially more profound in a fully turbulent flow.
INTRODUCTION
An essential aspect of virtually all astrophysical magnetic dynamos is the role played by turbulent magnetic field diffusion. From the analytical perspective, discussions of turbulent diffusion have until recently been generally couched in the language of mean field theory, and in particular, within a kinematic context (cf. Moffatt 1978; Krause & Rädler 1980). Indeed, the great theoretical elegance of mean field electrodynamics, together with its attractive intuitiveness, have led to a situation where basic constructs of this theory, such as turbulent diffusion and the ‘α-effect’, have carried over into domains, such as numerical simulations, where their meaningfulness is not a priori obvious (cf. Glatzmaier 1985). In a recent series of papers, we have examined precisely the question of how such notions can be carried over into the nonlinear domain, and further have asked under what circumstances nonlinear effects are likely to matter (Vainshtein & Rosner 1991; Cattaneo & Vainshtein 1991; Vainshtein & Cattaneo 1992; Tao, Cattaneo & Vainshtein 1993).
Ambipolar diffusion, or ion-neutral drift, has important effects on the transport of magnetic fields in weakly ionized media such as the galactic interstellar medium. Ambipolar diffusion can inhibit the development of small scale magnetic structure because the field ceases to be kinematic with respect to the ions at strengths well below equipartition with the neutrals. On the other hand, magnetic nulls are characterized by steep profiles in which the current density diverges. The addition of ambipolar diffusion to mean field α-ω dynamos makes the equations nonlinear and can lead to steady states or traveling waves.
INTRODUCTION
The theory of linear, kinematic, mean field dynamos has been studied extensively since the pioneering paper by Parker (1955). In such dynamos, the mean magnetic field grows despite the action of resistivity through the combined action of small-scale, helical motions (α effect) and large-scale shear flows (ω effect). If the background state is time independent, the mean field evolves exponentially in time, and saturation of the field amplitude must occur through effects not included in the model.
Astrophysical systems typically have very low resistivities and correspondingly high magnetic Reynolds numbers Rm (of order 108–1010 in the Solar convection zone and 1018–1020 in the galactic disk). This raises a problem for dynamo theory: if the resistivity is assumed to be molecular, the fastest growing wavelengths are extremely short and it is difficult to see how large scale fields could be generated. Moreover, the resistivity plays a central role in the calculation of the a effect (e.g., Moffatt 1978). Most workers therefore assume that turbulent resistivity is present.
The recent evidence for the possibility of laterally varying electrical conductivity in the lowermost mantle of the Earth has motivated us to consider in more detail the problem of dynamo action induced by this kind of inhomogeneity. An earlier model (Busse & Wicht 1992) has been extended in that the assumption of a thin layer of sinusoidal varying conductivity is replaced by the assumption of a thick layer. In the new formulation the toroidal field as well as the poloidal field are determined explicitly in the domain of varying conductivity. The results support the conclusion based on the earlier thin layer assumption that the dynamo action is too weak to be of geophysical importance.
INTRODUCTION
The influence of varying conductivity on the dynamo process has been investigated for example for galaxies (Donner & Brandenburg 1990) and accretion disks (Stepinski & Levy 1991) and found to be negligible there. Jeanloz's (1990) interpretation of the D′′ layer as a laterally inhomogeneous distribution of conducting and insulating alloys, resulting from chemical reactions at the core-mantle boundary and the percolation of iron into the mantle, has motivated us to consider the possibility of a dynamo induced by varying conductivity on the Earth's dynamo. Two questions arise in this context. Firstly, one may ask how a lower mantle with laterally varying conductivity will affect the extrapolation of magnetic fields from the Earth surface to the core. Poirier & le Mouel (1992) have investigated this question in detail and found the effect to be negligible. Jeanloz's (1990) view of pinned fieldlines is too dramatic.