You are on page 1of 53

Mixing and Dispersion in Flows

Dominated by Rotation and Buoyancy


1st Edition Herman J.H. Clercx
Visit to download the full and correct content document:
https://textbookfull.com/product/mixing-and-dispersion-in-flows-dominated-by-rotation
-and-buoyancy-1st-edition-herman-j-h-clercx/
More products digital (pdf, epub, mobi) instant
download maybe you interests ...

Turbulence and Dispersion in the Planetary Boundary


Layer 1st Edition Francesco Tampieri (Auth.)

https://textbookfull.com/product/turbulence-and-dispersion-in-
the-planetary-boundary-layer-1st-edition-francesco-tampieri-auth/

Context and Communication Herman Cappelen

https://textbookfull.com/product/context-and-communication-
herman-cappelen/

Mixing Music 1st Edition Russ Hepworth-Sawyer

https://textbookfull.com/product/mixing-music-1st-edition-russ-
hepworth-sawyer/

Biomotors: Linear, Rotation, and Revolution Motion


Mechanisms 1st Edition Peixuan Guo

https://textbookfull.com/product/biomotors-linear-rotation-and-
revolution-motion-mechanisms-1st-edition-peixuan-guo/
Markov Chains and Mixing Times Second Edition David A.
Levin

https://textbookfull.com/product/markov-chains-and-mixing-times-
second-edition-david-a-levin/

Dispersion Relations in Heavily-Doped Nanostructures


1st Edition Kamakhya Prasad Ghatak (Auth.)

https://textbookfull.com/product/dispersion-relations-in-heavily-
doped-nanostructures-1st-edition-kamakhya-prasad-ghatak-auth/

Visualization and Simulation of Complex Flows in


Biomedical Engineering 1st Edition Diego Gallo

https://textbookfull.com/product/visualization-and-simulation-of-
complex-flows-in-biomedical-engineering-1st-edition-diego-gallo/

Geography and the Political Imaginary in the Novels of


Toni Morrison 1st Edition Herman Beavers

https://textbookfull.com/product/geography-and-the-political-
imaginary-in-the-novels-of-toni-morrison-1st-edition-herman-
beavers/

Design and Development of Optical Dispersion


Characterization Systems Iraj Sadegh Amiri

https://textbookfull.com/product/design-and-development-of-
optical-dispersion-characterization-systems-iraj-sadegh-amiri/
CISM International Centre for Mechanical Sciences 580
Courses and Lectures

Herman J.H. Clercx


GertJan F. Van Heijst Editors

Mixing and Dispersion


in Flows Dominated by
Rotation and Buoyancy

International Centre
for Mechanical Sciences
CISM International Centre for Mechanical
Sciences

Courses and Lectures

Volume 580

Series editors

The Rectors
Elisabeth Guazzelli, Marseille, France
Franz G. Rammerstorfer, Vienna, Austria
Wolfgang A. Wall, Munich, Germany

The Secretary General


Bernhard Schrefler, Padua, Italy

Executive Editor
Paolo Serafini, Udine, Italy
The series presents lecture notes, monographs, edited works and proceedings in the
field of Mechanics, Engineering, Computer Science and Applied Mathematics.
Purpose of the series is to make known in the international scientific and technical
community results obtained in some of the activities organized by CISM, the
International Centre for Mechanical Sciences.

More information about this series at http://www.springer.com/series/76


Herman J.H. Clercx GertJan F. Van Heijst

Editors

Mixing and Dispersion


in Flows Dominated
by Rotation and Buoyancy

123
Editors
Herman J.H. Clercx GertJan F. Van Heijst
Department of Applied Physics Department of Applied Physics
Eindhoven University of Technology Eindhoven University of Technology
Eindhoven, Noord-Brabant Eindhoven, Noord-Brabant
The Netherlands The Netherlands

ISSN 0254-1971 ISSN 2309-3706 (electronic)


CISM International Centre for Mechanical Sciences
ISBN 978-3-319-66886-4 ISBN 978-3-319-66887-1 (eBook)
https://doi.org/10.1007/978-3-319-66887-1
Library of Congress Control Number: 2017951146

© CISM International Centre for Mechanical Sciences 2018


This work is subject to copyright. All rights are reserved by the Publisher, whether the whole or part
of the material is concerned, specifically the rights of translation, reprinting, reuse of illustrations,
recitation, broadcasting, reproduction on microfilms or in any other physical way, and transmission
or information storage and retrieval, electronic adaptation, computer software, or by similar or dissimilar
methodology now known or hereafter developed.
The use of general descriptive names, registered names, trademarks, service marks, etc. in this
publication does not imply, even in the absence of a specific statement, that such names are exempt from
the relevant protective laws and regulations and therefore free for general use.
The publisher, the authors and the editors are safe to assume that the advice and information in this
book are believed to be true and accurate at the date of publication. Neither the publisher nor the
authors or the editors give a warranty, express or implied, with respect to the material contained herein or
for any errors or omissions that may have been made. The publisher remains neutral with regard to
jurisdictional claims in published maps and institutional affiliations.

Printed on acid-free paper

This Springer imprint is published by Springer Nature


The registered company is Springer International Publishing AG
The registered company address is: Gewerbestrasse 11, 6330 Cham, Switzerland
Preface

The chapters contained in this volume correspond to the lectures given during the
course “Mixing and Dispersion in Flows Dominated by Rotation and Buoyancy”
that was held at the CISM in Udine (Italy), July 6–10, 2015.
Rotation and buoyancy play an essential role in many astrophysical, geophysical,
environmental, and industrial flows. They influence the transition to turbulence,
strongly affect large-scale (turbulent) flow properties by inducing anisotropy, and
also affect boundary-layer dynamics and inertial-range turbulence characteristics.
Moreover, rotation and buoyancy may have a strong impact on mixing, and on the
dispersion of passive and active tracers and of (inertial) particles and droplets in
such flows. The impact of buoyancy or rotation on transport may be direct (gravitational,
centrifugal, or Coriolis forces on fluid parcels or particles/droplets) or indirect by
the modified flow characteristics. These processes have direct relevance for heat
and mass transfer in many natural systems. Examples are (large-scale) convection
processes, transport of sediment in coastal flows, dispersion of suspended particulate
matter in estuarine flows, in lakes and reservoirs, and dispersion of aerosols and
pollutants in the atmospheric boundary layer. Increasing computational capabilities
and the rapid development of advanced experimental measurement tools, for
example optical diagnostics and particle tracking techniques, provide highly
resolved temporal and spatial data sets. This allows the exploration and analysis of
more complex flow phenomena and the associated transport processes in more
depth.
The aim of this course had been to present an overview of recent developments
in this field in a way accessible to participants coming from a variety of fields,
including turbulence research, (environmental) fluid mechanics, lake hydrody-
namics, or atmospheric physics. Topics to be discussed during the lectures ranged
from the fundamentals of rotating and stratified flows, mixing and transport in
stratified or rotating turbulence, transport in the atmospheric boundary layer, the
dynamics of gravity and turbidity currents, mixing in (stratified) lakes, and the
Lagrangian approach to analyze transport processes in geophysical and environ-
mental flows. This goal is not only reflected in this volume and the contributions are
aimed at doctoral students and postdoctoral researchers, but also at academic and

v
vi Preface

industrial researchers and practicing engineers, with a background in mechanical


engineering, applied physics, civil engineering, applied mathematics, meteorology,
physical oceanography, or physical limnology.
This volume starts with a general introduction on effects of rotation and density
stratification on fluid flows by GertJan van Heijst, followed by an overview of the
basic phenomena of turbulence and mixing in flows dominated by buoyancy by
Paul Linden. Subsequently, the connection is made with certain environmental flow
situations, such as mixing in lakes and reservoirs by Damien Bouffard and Alfred
Wuëst and energy balances in stably stratified, wall-bounded turbulence by Oscar
Flores and James Riley. The Lagrangian approach to unravel transport in turbulence
is addressed by Mickaël Bourgoin. Eckart Meiburg and Mohamad Nasr-Azadani
give an overview of recent numerical and modeling developments in the field of
gravity and turbidity currents. Finally, several aspects of transport processes in
rotating turbulence are reviewed by Herman Clercx.
The authors wish to thank all the contributors and the members of CISM for their
support and thoughtful suggestions during the course and the preparation of this
volume.

Eindhoven, The Netherlands Herman J.H. Clercx


GertJan F. Van Heijst
Contents

Effects of Rotation and Stratification: An Introduction . . . . . . . . . . . . . . 1


GertJan F. van Heijst
Turbulence and Mixing in Flows Dominated by Buoyancy . . . . . . . . . . . 25
Paul F. Linden
Mixing in Stratified Lakes and Reservoirs . . . . . . . . . . . . . . . . . . . . . . . . 61
Damien Bouffard and Alfred Wüest
Energy Balance in Stably-Stratified, Wall-Bounded Turbulence . . . . . . . 89
Oscar Flores and James J. Riley
Some Aspects of Lagrangian Dynamics of Turbulence . . . . . . . . . . . . . . 101
Mickaël Bourgoin
Gravity and Turbidity Currents: Numerical Simulations
and Theoretical Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 129
Eckart Meiburg and Mohamad M. Nasr-Azadani
Transport Phenomena in Rotating Turbulence . . . . . . . . . . . . . . . . . . . . . 181
Herman J.H. Clercx

vii
Contributors

Damien Bouffard Physics of Aquatic Systems Laboratory, Ecole Polytechnique


Federale de Lausanne, Lausanne, Switzerland
Mickaël Bourgoin Université de Lyon, ENS Lyon, Université Claude Bernard –
Lyon 1, CNRS, Laboratoire de Physique, Lyon, France
Herman J.H. Clercx Fluid Dynamics Laboratory, Department of Applied
Physics, Eindhoven University of Technology, Eindhoven, The Netherlands; J.M.
Burgers Center for Fluid Dynamics, Enschede, The Netherlands
Oscar Flores Departamento de Bioingeniería e Ingeniería Aeroespacial,
Universidad Carlos III de Madrid, Madrid, Spain
GertJan F. van Heijst Fluid Dynamics Laboratory, Department of Applied
Physics, Eindhoven University of Technology, Eindhoven, The Netherlands; J.M.
Burgers Center for Fluid Dynamics, Enschede, The Netherlands
Paul F. Linden Department of Applied Mathematics and Theoretical Physics,
Centre for Mathematical Sciences, University of Cambridge, Cambridge, UK
Eckart Meiburg Department of Mechanical Engineering, University of California
at Santa Barbara, Santa Barbara, CA, USA
Mohamad M. Nasr-Azadani Department of Mechanical Engineering, University
of California at Santa Barbara, Santa Barbara, CA, USA
James J. Riley Department of Mechanical Engineering, University of
Washington, Seattle, WA, USA
Alfred Wüest Swiss Federal Institute of Aquatic Science and Technology,
Kastanienbaum, Switzerland

ix
Effects of Rotation and Stratification:
An Introduction

GertJan F. van Heijst

Abstract Large-scale flows in the natural environment can be influenced by the


planetary rotation and also by density differences. This chapter aims to provide an
informal introduction into the effects of background rotation and stratification.

1 Effects of Rotation

The relative motion of a fluid in a rotating system is conveniently described by using


a co-rotating reference frame. Adopting such a reference system implies the appear-
ance of two additional forces in Newtons second law: the Coriolis and the centrifugal
forces. Expressed in a co-rotating coordinate frame the momentum equation takes
the following form:
Dv 1
+ 2 × v = − ∇P + ν∇ 2 v , (1)
Dt ρ

with v the relative velocity,  the rotation of the system, ρ the density, P the ‘reduced’
pressure, t the time, ν the kinematic fluid viscosity, and Dt D
≡ ∂t∂ + v · ∇ the mate-
rial derivative. The Coriolis acceleration is −2 × v, while the reduced pressure is
defined as
1
P = p + ρgrav − 2 r 2 , (2)
2
where p is the fluid pressure, grav the gravitational potential,  = || the angular
velocity, and 21 2 r 2 represents the centrifugal contribution (with r the distance mea-
sured with respect to the rotation axis). Note that the Coriolis acceleration −2 × v
is directed perpendicular to the fluid velocity v. On the northern hemisphere the

G.J.F. van Heijst (B)


Fluid Dynamics Laboratory, Department of Applied Physics, Eindhoven University
of Technology, P.O. Box 513, 5600MB Eindhoven, The Netherlands
e-mail: g.j.f.v.heijst@tue.nl
G.F. van Heijst
J.M. Burgers Center for Fluid Dynamics, Enschede, The Netherlands

© CISM International Centre for Mechanical Sciences 2018 1


H.J.H. Clercx and G.J.F. van Heijst (eds.), Mixing and Dispersion in Flows Dominated
by Rotation and Buoyancy, CISM International Centre for Mechanical Sciences 580,
https://doi.org/10.1007/978-3-319-66887-1_1
2 G.J.F. van Heijst

Coriolis induces a deflection to the right, while on the southern hemisphere a mov-
ing fluid experiences a deflecting force to its left.
By using a typical length scale L, a typical velocity U , and 1/ as a time scale,
the physical quantities may be non-dimensionalised as:

r = Lr̃ , v = U ṽ , t = t̃/ , P = ρU L p̃ (3)

with the tilde (∼) denoting nondimensionality. Equation (1) is then written in non-
dimensional form:

∂ ṽ ˜ + 2k × ṽ = −∇
˜ p̃ + E ∇˜ 2 ṽ
+ Ro(ṽ · ∇)ṽ (4)
∂ t̃

˜ the non-dimensional gradient operator.


with k = /|| the unit rotation vector and ∇
In this equation Ro and E represent the Rossby number and the Ekman number,
respectively, defined as
U [(v · ∇)v]
Ro ≡ ∼ , (5)
L [ × v]

ν [ν∇ 2 v]
E≡ ∼ . (6)
L 2 [ × v]

These nondimensional numbers indicate the relative importance of the advective


and viscous terms, respectively, compared to the Coriolis term in the momentum
equation. For a large-scale flow in the Earth’s atmosphere, with U ∼ 10 ms-1 and
L ∼ 1000 km, with  = 1 rev day-1 ∼ 10−4 s-1 , the Rossby number is typically Ro ∼
0.1. For oceanic flows, with U ∼ 1 ms-1 , on such a scale, Ro ∼ 10-2 . Apparently, the
Coriolis force plays a nearly dominant role in such flows. Such large-scale flows are
turbulent, implying that the kinematic viscosity ν should be replaced by a turbulent
eddy viscosity νt . Nevertheless, the Ekman number for such large-scale flows is
usually much less than unity, implying that viscous effects are only of very marginal
importance (as long as the gradients in the flow field are not too large). For a steady
flow with Ro  1 and E  1, the momentum Eq. (4) reduces to

2k × v = −∇p (7)

(note that from hereon the tildes have been dropped). This expression reveals a
balance between the Coriolis force and the pressure gradient, which is referred to as
the geostrophic balance. Apparently, the velocity v is directed perpendicular to the
pressure gradient, in other words: the flow is directed along isobars. A remarkable
property of geostrophically balanced flow is derived by taking the curl of Eq. (7),
yielding
∂v
(k · ∇)v = 0 → =0. (8)
∂z
Effects of Rotation and Stratification: An Introduction 3

This so-called Taylor-Proudman theorem expresses that the geostrophic flow is


independent of the axial coordinate (z), see Proudman (1916). The validity of this
theorem was demonstrated experimentally by Taylor (1917) by slowly towing a
cylindrical obstacle through a rotating fluid: the obstacle carried an axially aligned
column of stagnant fluid, while moving steadily moving through the rotating fluid
tank. Although Eq. (8) is strictly only valid for Ro  1, such columnar structures
are common features of rotating flows, even under conditions Ro ∼ 1. They are
commonly referred to as ‘Taylor columns’.

1.1 Flow on a Spherical Planet

The oceans and atmosphere on our planet can be considered as relatively thin fluid
shells on a rotating solid sphere. Due to this geometry, the large-scale oceanic and
atmospheric flows can be considered as motions in the locally horizontal plane. This
implies that the local vorticity vector ω is directed vertically upwards. For flow not
too close to the poles, it is convenient to adopt a Cartesian coordinate system (x, y, z),
with the coordinates pointing in eastward, northward, and vertically upward direction,
respectively (see Fig. 1). The corresponding velocity components are v = (u, v, w).
Decomposing the rotation vector in these coordinate directions gives

 = (x , y , z ) = (0,  cos φ,  sin φ) , (9)

with φ the geographical latitude. The x, y, z-components of the Coriolis term in


Eq. (1) are then easily derived:

2 × v = 2(w cos φ − v sin φ, u sin φ, −u cos φ) . (10)

Fig. 1 Definition sketch of


coordinates on a rotating
sphere
y
z
4 G.J.F. van Heijst

The thin-shell approximation implies w  u, v, so that expression (10) can be


approximated by
2 × v = (−f v, +fu, −2u cos φ) (11)

with
f = 2 sin φ (12)

the so-called Coriolis parameter. The x, y, z-components of Eq. (1) can then be writ-
ten as
Du uv tan φ uw 1 ∂p
− + − 2v sin φ + 2w cos φ = − + Fx , (13a)
Dt R R ρ ∂x
Dv u2 tan φ vw 1 ∂p
+ + + 2u sin φ = − + Fy , (13b)
Dt R R ρ ∂y
Dw u2 + v 2 1 ∂p
− − 2u cos φ = − − g + Fz , (13c)
Dt R ρ ∂z

with R the planet’s radius and Fx , Fy , Fz representing e.g. viscous forces. For the
large-scale flows in the Earth’s atmosphere and oceans, a scaling analysis reveals
that the curvature-related terms in Eq. (13) are small compared to the others, in such
a way that the equations (for the inviscid case) take on the following form:

∂u ∂u ∂u 1 ∂p
+u +v −fv = − , (14a)
∂t ∂x ∂y ρ ∂x
∂v ∂v ∂v 1 ∂p
+u +v + fu = − , (14b)
∂t ∂x ∂y ρ ∂y
1 ∂p
0=− −g, (14c)
ρ ∂z

with f the Coriolis parameter as defined in Eq. (12). Apparently, curvature effects
associated with the Earth’s spherical shape are negligible, except in the Coriolis
terms.
The Coriolis parameter is a nonlinear function of the geographical latitude φ. For
limited latitudinal scales (L < 1000 km) we may assume f to be constant:

f = f0 = 2 sin φ0 , (15)

with φ0 a reference latitude. This is the so-called f -plane approximation, which


ignores the latitudinal variation of f . In a better approximation, one may expand
f (φ) in a Taylor series:

cos φ0
f (φ) = f (φ0 + δφ) = 2[sin φ0 + Rδφ + O(δφ2 )] , (16)
R
Effects of Rotation and Stratification: An Introduction 5

where Rδφ = y, the northward coordinate. For small δφ-values, the higher-order
terms in the expression are negligibly small, leaving:

2 cos φ0
f = f0 + βy, with β = . (17)
R
This linear approximation of f (φ) is the so-called β-plane approximation, which
is relevant for large-scale flows with a moderate latitudinal extent. In particular, the
β-plane approximation is relevant for flows in the equatorial region, for which f0 = 0.

1.2 Some Horizontal Balances

Vortical motions are common features of large-scale flows, both in the atmosphere
and in the ocean. Examples are the high- and low-pressure cells in the atmosphere
and the vortices shed from ocean currents such as the Gulf Stream. It is instructive
to consider a model curvilinear flow around a common centre, for which centripetal
accelerations may be important. For steady inviscid flow with velocity V around a
circular trajectory with curvature radius R, the basic (dimensional) balance is written
as
V2 1 ∂p
+fV = , (18)
R ρ ∂r

with r the local radial coordinate. In this expression one recognizes the nonlinear
centrifugal acceleration, the Coriolis acceleration, and the radial pressure gradient.
When written in nondimensional terms, the Rossby number Ro would appear as a
prefactor of the centrifugal term. Depending on the Ro-value, different balances are
possible, see e.g. Holton (1992).
Ro  1 : geostrophic balance

1 ∂p
fV = . (19)
ρ ∂r

When the centrifugal term is negligibly small, the flow is governed by a balance
between the pressure gradient force and the Coriolis force. A low-pressure centre
(∂p/∂r > 0) corresponds with a cyclonic flow, and vice versa a high-pressure centre
corresponds with an anticyclonic circulation.
Ro  1 : cyclostrophic balance

V2 1 ∂p
= . (20)
R ρ ∂r

For Ro  1, the centrifugal term dominates the Coriolis term, such that the flow is
governed by a balance between the pressure-gradient force and the centrifugal force.
6 G.J.F. van Heijst

As Eq. (20) reveals, cyclostrophic motion is in all cases around a low-pressure cell
(∂p/∂r > 0). Moreover the flow can be either cyclonic or anticyclonic. This regime
corresponds to e.g. tornadoes, but also to the bathtub vortex.
Ro ∼ O(1) : gradient flow

V2 1 ∂p
+fV = . (21)
R ρ ∂r

For a given pressure gradient the solution for the velocity V is:

1 1 R ∂p 1/2
V = − fR ± [ f 2 R2 + ] . (22)
2 4 ρ ∂r

Four cases are possible, depending on the sign of the pressure gradient ∂p/∂r.
However, two of these cases are not stable. The two ‘regular’ cases correspond to
a ‘regular low’, with cyclonic motion (V > 0) governed by a balance between the
outward centrifugal and Coriolis forces and the inward pointing pressure gradient
force. In the other case, the ‘regular high’, the motion is anticyclonic (V < 0) and
governed by a balance between the inward Coriolis force and the outward centrifugal
and pressure gradient forces.
inertial motion
The balance Eq. (18) for steady motion along circular trajectories also describes a
balanced flow for the case of a zero radial pressure gradient. In that case the equation
is reduced to
V2
+fV = 0. (23)
R
Apart from the trivial solution V = 0 this balance between centrifugal and Coriolis
forces allows inertial motion, with V = −fR = constant. Apparently, this inertial
flow corresponds with fluid motion along circular paths in anticyclonic direction.
Alternatively, the motion may also be described in Cartesian coordinates, namely by

du dv
−fv = 0 , + fu = 0 , (24)
dt dt
with solutions
u(t) = V cos ft, v(t) = V sin ft . (25)

This solution describes harmonic motion along circular paths with frequency f ,
i.e. with inertial period T = 2π/f . This type of motion is typically observed in large
lakes or semi-enclosed seas like the Baltic after some disturbance (such as the passage
of an atmospheric low-pressure cell) has occurred.
Effects of Rotation and Stratification: An Introduction 7

1.3 How to Create Vortices in the Lab

Vortex structures may be created in a rotating fluid confined in a container mounted


on a turntable in a number of different ways, see e.g. van Heijst and Clercx (2009).
One way is by locally stirring the fluid confined in a smaller bottomless cylinder that
is placed vertically in the otherwise solidly rotating fluid. By vertically lifting the
cylinder out of the fluid the ‘stirring vortex’ is released. After some adjustment this
vortical flow becomes nicely organised in the form of an axially aligned columnar
structure. In principle one may thus create both cyclonic and anticyclonic vortices
in this way. It was found, however, that the anticyclonic vortices are very unstable:
they usually fall apart, eventually resulting in two dipolar structures, see Kloosterziel
and van Heijst (1991). An alternative generation method is by having the fluid level
inside the generating cylinder initially somewhat lower than outside: upon lifting the
cylinder, the ‘gravitational collapse’ results in a radially inward motion, which—by
conservation of angular momentum—is quickly converted into a cyclonic vortex flow.
Another generation method is by syphoning fluid through a perforated tube placed
vertically in the rotating fluid tank. Again, by conservation of angular momentum
the radial flow thus induced is converted into a cyclonic swirl motion around the
perforated tube. After carefully lifting the the tube out of the fluid, a stable cyclonic
‘sink vortex’ is produced.
The vorticity and swirl velocity profiles of the ‘sink’ and ‘stirring’ vortices are
very well fitted by

r2
ωsink (r) = ω0 exp(− ), (26a)
R2
ω0 R 2
r2
vsink = [1 − exp(− 2 )] , (26b)
2r R
and

r2 r2
ωstir (r) = ω0 (1 − ) exp(− ), (27a)
R2 R2
1 r2
vstir = ω0 r exp(− 2 ) , (27b)
2 R
with ω0 a vorticity amplitude and R some radial length scale (see Kloosterziel and
Heijst 1992).
These generation techniques have been applied in a variety of experimental stud-
ies of barotropic vortices, see e.g. Kloosterziel and van Heijst (1991), Trieling et al.
(1998), and Zavala Sansón et al. (2012). Similar techniques may be applied to gener-
ate baroclinic two-layer vortices. Additionally, anticyclonic two-layer vortex struc-
tures can be created by adding a source of dense fluid (at the bottom) or a source of
lighter fluid (at the free surface): the radial spreading of the source fluid results in an
8 G.J.F. van Heijst

anticyclonic motion of the injected fluid. After removal of the source, an anticyclonic
two-layer vortex has been established.

1.4 The Ekman Layer

Although for E  1 viscous effects play a minor role in most of the flow domain, near
solid domain boundaries—where velocity gradients become large—viscous effects
are essential in order to let the geostrophic flow in the bulk of the domain satisfy the
boundary conditions (no-slip at solid walls and stress conditions at the free surface).
We will here confine ourselves to the Ekman layer, which is the boundary layer at a
horizontal (or weakly sloping) solid boundary.
For a steady, small-Ro flow the momentum Eq. (4) becomes (after dropping the
tildes)
2k × v = −∇p + E∇ 2 v . (28)

Since E  1 the viscous term is generally very small, but it may become important
when the normal gradients at a boundary become large. At a horizontal boundary the
viscous term becomes equally important as the Coriolis term when

∂2 E
E∇ 2 ∼ E ∼ 2 ∼ O(1) , (29)
∂z2 δ

from which one derives that the nondimensional thickness of the Ekman layer is
δ ∼ E 1/2 and in dimensional form:
ν 1/2
δE = Lδ = LE 1/2 = ( ) . (30)

Note that the (constant) thickness of the Ekman layer is independent of the size
of the flow configuration. In a laboratory experiment in a water-filled rotating tank
with ν = 10−6 m2 s−1 and  = 10−4 s−1 (the Earth makes one revolution in 24 h),
one derives δE  30 m.
The Ekman layer over a solid horizontal boundary serves to provide the matching
of the geostrophic flow in the ‘interior’ domain to the no-slip boundary condition.
If the horizontal velocity in the interior is vI = (uI , vI ) = O(1), with the subscript I
referring to ‘interior’ and u and v the velocity components in a corotating Cartesian
x, y-frame, this requires that the velocity in the Ekman layer (subscript E) is vE
= (uE , vE ) = O(1). As the Ekman layer is very thin (since E  1), this implies
pI ∼ pE = O(1). The x, y-components of Eq. (28) can then be written as
Effects of Rotation and Stratification: An Introduction 9

∂pI ∂ 2 uE
−2vE = − +E 2 , (31a)
∂x ∂z
∂pI ∂ 2 vE
+2uE = − +E 2 , (31b)
∂y ∂z

while the geostrophic flow in the interior regime is governed by the geostrophic
balance (7), or
∂pI ∂pI
− 2vI = − , + 2uI = − . (32)
∂x ∂y

With the bottom at z = 0, the boundary conditions for the Ekman-layer solution
of the Eqs. (31a) and (31b) are

z = 0 : uE = vE = 0 (no slip) , (33a)


z → ∞ : (uE , vE ) → (uI , vI ) . (33b)

The solution to this problem is

uE = uI − [uI cos(z/δ) + vI sin(z/δ)]e−z/δ , (34a)


vE = vI − [vI cos(z/δ) − uI sin(z/δ)]e−z/δ . (34b)

These solutions show an oscillating behaviour in the z-direction, although damped.


Moreover, the horizontal flow in the Ekman layer has a significant component also
in the direction perpendicular to the principal direction of the geostrophic flow. This
can be nicely seen by taking vI = (uI , 0), for which Eq. (34) become

uE = uI [1 − e−z/δ cos(z/δ)] , (35a)


vE = uI e−z/δ sin(z/δ) . (35b)

Apparently, the Ekman-layer flow contains a y-component, while the geo-strophic


flow is purely in the x-direction. By integrating the continuity equation ∇ · v = 0
for incompressible flow over the thickness of the Ekman layer, with the condition
wE (z = 0) = 0, one arrives at

1 1/2
wE (z > δ) = E ωI , (36)
2
with ωI the z-component of the relative vorticity of the geostrophic interior flow. This
is the so-called ‘Ekman suction condition’, which the Ekman layer imposes on the
geostrophic flow outside the viscous layer. A non-zero suction velocity is directly
associated with a horizontal divergence of the flow in the Ekman layer. By integrating
the Ekman contributions (uE − uI ) and (vE − vI ) over the thickness of the Ekman
layer, one derives the following expressions of the Ekman volume fluxes in the x-
and y-directions, respectively:
10 G.J.F. van Heijst

1
Mx = − (uI + vI )E 1/2 , (37a)
2
1
My = − (vI − uI )E 1/2 . (37b)
2

Note that these are two-dimensional fluxes, with dimensions [m2 s−1 ], so fluxes
per unit length in the perpendicular direction.
For the special case vI = (uI , 0) the Ekman transport components are

1
Mx = −My = − uI E 1/2 , (38)
2
so the Ekman layer accommodates a mass flux in a direction perpendicular to that
of the principal flow in the interior.
Note that, since uE , vE ∼ O(1) and δ = E 1/2 , the horizontal Ekman fluxes are
O(E 1/2 ). Likewise, the suction velocity is wE ∼ O(E 1/2 ). Although this suction
velocity is small compared to the principal horizontal velocity components of the
geostrophic interior flow, this Ekman suction velocity may play a crucial role in the
development of the geostrophic interior flow. For example, in the case of a cyclonic
or anticyclonic vortex above a flat horizontal bottom the Ekman layer leads to a
‘blowing’ or ‘suction’ flow, which sets up an internal circulation in the vortex that
gradually changes its vorticity structure. From Eq. (36) one sees that positive vortic-
ity (ωI > 0, as in the core of a cyclonic vortex) implies wE (z > δ) > 0, i.e. Ekman
‘blowing’, while the opposite takes place for ωI < 0: Ekman ‘suction’. This mech-
anism is illustrated in Fig. 2.

Spin-up/spin-down Most vortices, such as the ‘stirring’ vortices produced in a


laboratory experiment, see (27), have a double-signed vorticity structure: a cyclonic
vortex core is shielded by a region of negative vorticity. In such a case the Ekman
layer produces ‘blowing’ under the vortex core, while ‘suction’ is produced under
the shielding outer band. As a result a (weak) internal circulation is generated with a
radial velocity (uI > 0) inside the vortex. Conservation of angular momentum (rvθ =
constant) implies that the swirl velocity of fluid elements brought to a larger radius

(a) (b)
f/2 f/2

>0 <0
z z

E E

Fig. 2 The bottom Ekman layer generates a ‘blowing’ in the core of a cyclonic vortex, and b
‘suction’ in the core of an anticyclonic vortex
Effects of Rotation and Stratification: An Introduction 11

decreases. Obviously, this internal circulation driven by the Ekman suction/blowing


results in the spin-down of cyclonic vortices. Likewise, anticyclonic vortices will
show spin-up by the same mechanism.
Exactly the same mechanism governs the spin-up of fluid in a rotating cylindrical
container. Immediately after increasing the container rotation from  to  + 
Ekman layers are formed at the bottom and top endplates of the cylinder. As the
fluid in the bulk lags behind, it has negative relative vorticity (with respect to the
coordinate system corotating with the container): ωI < 0. As a result Ekman suction
will occur at both Ekman layers, thus setting up an internal circulation with uI < 0.
The fluid elements thus brought to smaller radii will acquire an increasing azimuthal
velocity, until eventually all the fluid rotates with the new rotation speed  + .
The timescale associated with this adjustment process is called the ‘Ekman spin-up
timescale’, which is given by

H
TE = (E 1/2 )−1 = . (39)
(ν)1/2

This timescale also applies to more general flow adjustments (spin-up, spin-down)
in a fluid layer with depth H.

Wind-Driven Ekman layer Although the above analysis was confined to a


geostrophic flow over a solid boundary (such as wind blowing over a ground sur-
face, a sea current flowing over the sea bottom, barotropic vortices in a rotating
water-filled laboratory tank with a solid bottom), it may be extended to the case of
a stress exerted at the surface of the fluid. This corresponds to the case of a wind
stress exerted on the surface of the ocean. An Ekman layer will be formed at the
surface, which can be analysed in the same way as for the Ekman at a solid no-slip
boundary. Most strikingly, it is then found that the Ekman layer carries a net mass
flux perpendicular to the direction of the wind stress exerted at the ocean surface: to
its right on the northern hemisphere, and to its left on the southern hemisphere. This
feature is responsible for so-called ‘coastal upwelling’, when the prevailing winds
in a coastal area induce an Ekman transport directed away from the coast: by mass
conservation this flux has to be compensated by upwelling of deeper ocean water.

1.5 The Shallow-Water Approximation

Consider the flow in a layer of fluid over some fixed bottom in a rotating system.
For convenience, we adopt a Cartesian coordinate system, with z pointing vertically
upwards. Viscous effects are neglected. In view of the background rotation it is
assumed that the vertical velocity w is much smaller than the horizontal components,
i.e. u, v  w, and besides that the horizontal flow (u, v) is independent of z. In other
words: the fluid moves in the form of vertical columns. These columns may be
stretched or squeezed when moving over bottom topography, or due to elevation
changes of the free surface, but they do not tilt.
12 G.J.F. van Heijst

The flow is assumed to be in hydrostatic balance, so that the horizontal pressure


gradients are directly linked with the gradients in the free-surface elevation. Writing
out the x, y-components of the momentum equation leads, after cross-differentiation
and subtraction, to the following relation:

1 D ∂u ∂v
(f + ζ) + ( + )=0 (40)
f + ζ Dt ∂x ∂y

with f the Coriolis parameter and

∂v ∂u
ζ ≡ ωz = − (41)
∂x ∂y

the z-component of the relative vorticity. The relation (40) expresses that a horizon-
tal convergence ( ∂u
∂x
+ ∂v
∂y
< 0) leads to an increase in the vorticity (f + ζ), while
divergence results in a decreasing vorticity. This is directly related to stretching and
squeezing of vertical columns in the regions of convergence and divergence, respec-
tively. By integrating the continuity equation in the vertical one derives a relation
between the horizontal divergence and the change in the fluid depth:

1 DH ∂u ∂v
+( + )=0. (42)
H Dt ∂x ∂y

Combination of Eqs. (40) and (42) results in

D f +ζ
( )=0, (43)
Dt H

which expresses conservation of ‘potential vorticity’ (f + ζ)/H. This quantity plays


an essential role in geophysical fluid dynamics (GFD) theory, and its conservation
plays a key role in the understanding of many fundamental phenomena in this field.
For example, the north-south motion of a vortex column implies changes in its relative
vorticity (even in the case H = constant): according to Eq. (43) a cyclonic vortex (with
ζ > 0) moving northward will become weaker, while an anticyclonic vortex (with
ζ < 0) moving in the same direction will intensify. Furthermore, the property of
conservation of potential vorticity allows for a topographic modelling of the β-effect
in the laboratory. Whereas a fluid column (with constant H) in the GFD situation
when moving in north/south direction experiences changes in its relative vorticity ζ
due to changes in the Coriolis parameter f , a proper bottom topography in a rotating
tank (with uniform f = 2) may lead to similar changes in ζ when a fluid column
moves up or down the topographic slope. Of course, this is all under the assumption
of columnar fluid motion and that viscous effects are negligible.
Conservation of potential vorticity also explains the meandering motion of an
eastward flow over a topographic ridge. The mechanism works as follows: a fluid
column when moving up the mountain ridge is squeezed (H < H0 ) and thus acquires
Effects of Rotation and Stratification: An Introduction 13

negative relative vorticity (ζ < ζ0 = 0), implying a deflection of the flow in south-
ward direction. Once it has crossed the mountain ridge the column has obtained
its initial length H0 , but it has arrived at a more southern latitude, so that the local
f -value has decreased: f < f0 . As a consequence the relative vorticity of the flow is
then ζ > 0, implying a deflection to the ‘left’, i.e. in northern direction. The flow
(with constant depth H) then sets in a meandering motion in eastern direction, cen-
tred around its initial latitude φ0 , such that the total vorticity remains conserved. The
same mechanism is responsible for the existence of Rossby waves, by virtue of the
gradients in the background vorticity.

2 Density Effects

Effects of density differences and gradients play an essential role in environmental


and geophysical flows, both on smaller and larger scales. In the atmosphere these
density differences may be due to differences in temperature and humidity, while in
the ocean, seas, estuaries they are usually due to a combination of temperature and
salinity. Additionally, suspended material will also lead to net differences in density.

2.1 Density Currents

Horizontal density differences or gradients usually result in density currents or gravity


currents, in which the denser component flows underneath the lighter fluid, or vice
versa, in which the lighter fluid flows over the denser fluid. Examples of such density
currents are found in the salt-tongue intrusion in a river mouth, where salty sea water
flows inland over the river bed, and also in the flow of cold air over the floor of a
room when opening a window on a cold winter day. Such flows can be conveniently
modelled in the laboratory in a so-called ‘lock-exchange’ experiment as sketched in
Fig. 3.
In this experiment, a vertical barrier initially separates two fluids of different
densities ρ and ρ + ρ. When working with water, the density difference is conve-

Fig. 3 The lock-exchange


experiment

+
14 G.J.F. van Heijst

niently obtained by dissolving some amount of salt in one of the fluid compartments).
After removing the barrier, a bottom current of denser fluid and a surface current
of lighter fluid set in. The gravitational driving force of both currents is of course
essentially dependent on the density difference ρ. It is convenient to define the
so-called ‘reduced gravity’ g , which is defined as


g = g, (44)
ρ

where g is the magnitude of the gravitational acceleration. In most geophysical or


environmental situations ρ/ρ is very small, implying that g  g. For this reason
internal gravity currents such as in the lock-exchange experiment (Fig. 3) are much
slower than gravitational flows with ρ ≈ ρ. Many aspects of gravity currents have
been studied by many researchers, see e.g. the review by Simpson (1999). In the early
years research on gravity currents was mainly experimental and theoretical modeling
was based on the shallow-water approximation. In recent years more sophisticated
high-resolution numerical modeling studies became possible, in which even details
of the turbulent mixing could be taken into account. See, for example, the work
by Härtel et al. (2000).

2.2 Stability

Let us now consider the case of a vertical density stratification, as sketched in Fig. 4.
Such a configuration, which is easy to realise in the laboratory, can be consid-
ered as a model situation of a realistic density distribution, such as encountered
in the atmosphere or in oceans, seas and inland waters. In order to analyse pos-
sible fluid motion in this density configuration—which is assumed horizontally
homogeneous—we perform a hypothetical experiment: we consider a small fluid
parcel of density ρ(z0 ) at initial level z = z0 and displace it upwards over a small
distance ζ. The ambient density in its new position is then equal to ρ − δρ, with

Fig. 4 Schematic
configuration of a vertical
z (z)
density stratification 
g

z = z0

Effects of Rotation and Stratification: An Introduction 15


δρ = − ζ. (45)
dz

The displaced particle thus experiences a ‘restoring’ force (per unit mass) equal to
gδρ, which tends to bring the parcel back to its initial position. Such a stratification
is statically stable if dρ/dz < 0. If the density would decrease with depth, the dis-
placed particle would experience a buoyancy force in upward direction, and it would
accelerate vertically upwards. Such a stratification with dρ/dz > 0 is thus statically
unstable.
As remarked, the restoring gravity force acting on the displaced fluid parcel is
equal to gδρ, with δρ given by relation (45). As a result, the parcel will experience a
vertical acceleration d 2 ζ/dt 2 , and the equation of motion (excluding viscous effects)
is then
d2ζ dρ
ρ 2 =g ζ. (46)
dt dz

This equation can also be written as

d2ζ
+ N 2ζ = 0 , (47)
dt 2
with N the buoyancy frequency defined as

g dρ
N2 = − . (48)
ρ dz

For a statically stable stratification N is real, and the solution of the differential
Eq. (47) describes a harmonic oscillation around the equilibrium level ζ = 0: appar-
ently, an initial perturbation in a stably stratified fluid results in wavelike motions.
On the other hand, in the case of an unstable stratification the solution has a run-
away character: perturbations in an unstably stratified fluid are strongly amplified
and result in vigorous overturning motions, and hence mixing.

2.3 Baroclinic Generation of Vorticity

The motion of an incompressible inviscid stratified fluid is governed by the equa-


tions of momentum and mass conservation and by the incompressibility condition,
respectively:
Dv ∂v
ρ = ρ[ + (v · ∇)v] = −∇p + ρg , (49)
Dt ∂t

Dρ ∂ρ
ρ = + (v · ∇)ρ = 0 , (50)
Dt ∂t
16 G.J.F. van Heijst

(a) (b)

p Thigh river
Tlow
sea p

Fig. 5 Baroclinic vorticity generation in two different cases: a sea breeze, b salt intrusion in a river
mouth

∇·v =0. (51)

By taking the curl of the momentum equation one derives

Dω ∂ω 1
= + (v · ∇)ω = (ω · ∇)v + 2 ∇ρ × ∇p , (52)
Dt ∂t ρ

which is usually referred to as the vorticity equation. It describes the change of the
vorticity of a fluid element due to stretching/squeezing/tilting of vortex tubes (first
term on the r.h.s. of Eq. (52)) and due to baroclinic production (second term on the
r.h.s.). The latter effect is present whenever the gradients of density and pressure are
not aligned (i.e. ∇ρ × ∇p = 0). This situation is encountered, for example, in the
case of a sea breeze on a sunny day near the coast, and in the case of a salt intrusion
into a river mouth (see Fig. 5).

2.4 Boussinesq Approximation

In many geophysical situations the variations in density and pressure may be con-
sidered as small perturbations (ρ and p ) with respect to their static equilibrium
distributions ρ0 (z) and p0 (z), respectively:

ρ(x, t) = ρ0 (z) + ρ (x, t) , p(x, t) = p0 (z) + p (x, t) . (53)

Substitution into Eq. (50) yields

Dv
(ρ0 + ρ ) = −∇(p0 + p ) + (ρ0 + ρ )g . (54)
Dt
Under the assumption of a basic hydrostatic balance, i.e. 0 = −∇p0 + ρ0 g, this
equation may be written as
Effects of Rotation and Stratification: An Introduction 17

ρ Dv 1 ρ
(1 + ) = − ∇p + g , (55)
ρ0 Dt ρ0 ρ0

For ρ /ρ0  1, this equation becomes

Dv ∂v 1 ρ
= + (v · ∇)v = − ∇p + g . (56)
Dt ∂t ρ0 ρ0

In this so-called Boussinesq approximation the density variations only arise in


the gravitational term, which is usually referred to as the buoyancy term, which is
essentially the reduced gravity.

2.5 Waves in a Stratified Fluid

A stratified fluid may support internal waves, as already discussed in Sect. 2.2. In the
case of a discrete stratification, for example in the form of two fluid layers separated
by a sharp interface, the internal wave motion is easily observed in the motion of
the interface. Such internal waves can be analysed in the same fashion as gravity
waves at a free fluid surface. In the case of a continuous stratification the analysis
of the internal wave motion is more subtle, and one has to start from the Eqs. (49),
(50) and (51), expressing conservation of momentum, conservation of mass, and
the continuity condition, respectively. In the Boussinesq approximation this set of
equations becomes
∂v 1 ρ
+ (v · ∇)v = − ∇p + g , (57)
∂t ρ0 ρ0

∂ρ
+ (v · ∇)ρ = 0 , (58)
∂t
∇·v =0. (59)

When additionally restricting the analysis to wave motions with small amplitudes
one may linearise the equations of motion. Written in terms of a Cartesian coordinate
system (x, y, z), with z vertically upwards, and (u, v, w) the corresponding velocity
components, the linearised equations take the following form:

∂u 1 ∂p
=− , (60)
∂t ρ0 ∂x

∂v 1 ∂p
=− , (61)
∂t ρ0 ∂y

∂w 1 ∂p ρg
=− − , (62)
∂t ρ0 ∂z ρ0
18 G.J.F. van Heijst

∂ρ ρ0
− N2 w = 0 , (63)
∂t g

∂u ∂v ∂w
+ + =0. (64)
∂x ∂y ∂z

These are five equations for the five variables u, v, w, ρ and p . Elimination in
favour of w results in

∂2w 2 ∂ w
2
∂2w
∇2[ ] + N [ + ]=0, (65)
∂t 2 ∂x 2 ∂y2

with the buoyancy frequency defined as N 2 = −(g/ρ0 )dρ0 /dz. Solution of Eq. (65)
for arbitrary N(z) is far from trivial. In order to make the analysis feasible, we consider
the most simple case N = constant. For ρ  ρ0 (z0 ) this corresponds with a linear
stratification of the form ρ0 (z) = ρ̄{1 + α(z0 − z)}. We assume wave-like solutions
of the following form

w(x, y, z, t) = ŵ exp[i(kx + ly + mz − ωt)] , (66)

with ŵ the constant amplitude, ω the frequency, and K = (k, l, m) the wave num-
ber vector. Substitution of the general solution (66) into (65) yields the following
dispersion relationship:
(k 2 + l 2 )
ω2 = N 2 2 . (67)
(k + l2 + m2 )

For the most simple case K = (k, 0, 0) the solution reduces to

w(x, t) = ŵ cos(kx − ωt) , (68)

which represents waves with frequency ω = N propagating in the x-direction. The


wave period is T = 2π/ω = 2π/N, while the wavelength is λ = 2π/k. This is a
transversal wave: the fluid particles move vertically up and down with velocity w,
whereas the wave propagates in the x-direction at phase speed

ω λN
c= = . (69)
k 2π

For the general case with K = (k, l, m) the dispersion relationship (67) can also
be written as
ω (k 2 + l 2 )1/2
=± 2 = ± cos θ , (70)
N (k + l2 + m2 )1/2

where θ is the angle between the lines of constant phase and the vertical. Some of
the wave characteristics are sketched in Fig. 6.
Effects of Rotation and Stratification: An Introduction 19

cg

x
Fig. 6 Schematic illustration of some internal wave characteristics

cg cg

c c

g
(z)
c c

cg cg

Fig. 7 Internal waves generated by an oscillating cylinder

The discussion can be simplified by choosing the x-axis such that the wave-
number vector K lies in the xz-plane (this can be done without loss of generality), so
that in this new coordinate frame we have K = (k, 0, m). The dispersion relationship
(70) then takes the following form:
ω k
=± = ± cos θ . (71)
N |K|

Waves of this type can be generated in the laboratory by a horizontal cylinder


performing vertical oscillatory motions (with frequency ω) in a linearly stratified
fluid, see Fig. 7. The internal wave motion, visualised by the shadowgraph technique,
is confined to an X-shaped pattern, the legs making an angle θ with the vertical.
20 G.J.F. van Heijst

2.6 Vortices in a Stratified Fluid

Vortex structures are observed both in the atmosphere and in the oceans. These
vortices occur on a range of length scales, and often show complicated interactions.
In many cases stratification effects play an important role in the dynamical evolution
of the vortex structures.

Generation of vortices in the laboratory Vortices in a stratified fluid may be


generated in the lab in a number of different ways, some of which are schematically
drawn in Fig. 8. Vortices are easily produced by localized stirring with a rotating,
bent rod or by using a spinning sphere. In both cases the rotation of the device adds
angular momentum to the fluid, which is swept outwards by centrifugal forces. After
some time the rotation of the device is stopped, upon which it is lifted carefully out
of the fluid. After a short while the turbulence introduced during the forcing has
decayed, and a laminar horizontal vortex motion results. The shadowgraph visual-
izations shown in Fig. 9 clearly reveal the turbulent region during the forcing by the
spinning sphere and a more smooth density structure soon after the forcing is stopped.
Vortices produced in this way typically have a pancake’ shape, with the vertical size
of the swirling fluid region being much smaller than its horizontal size. This implies
large gradients of the flow in the vertical direction and hence the presence of a radial
vorticity component ωr . Although the swirling motion in these thin vortices is in
good approximation planar, the significant vertical gradients imply that the vortex
motion is not 2D. Additionally, the strong gradients in z-direction imply a significant
effect of diffusion of vorticity in that direction.
Alternatively, a vortex may be generated by tangential injection of fluid in a thin-
walled, bottomless cylinder, as also shown in Fig. 9. The swirling fluid volume is
released by lifting the cylinder vertically. After some adjustment, again a pancake-
like vortex is observed with features quite similar to the vortices produced with the
spinning devices.

Fig. 8 Generation techniques of vortices in a stratified fluid, see Flór (1994)


Another random document with
no related content on Scribd:
upon them a ray which they will reflect and send back to his own
eye, he will see in the centre of each a spot with the brightness of
burnished silver. He will be tempted to think that this spot is a bubble
of air; but, by immersing the ice in hot water, you can melt the ice all
around the spot,—and when it alone remains, you will see it diminish
and disappear without any trace of air. The spot is a vacuum. Such is
the faithfulness to herself with which Nature operates; thus, in all her
operations, does she submit to her own laws. We know that ice, in
melting, contracts; and here we arrest the contraction, as it were, in
the very act. The water of the flowers cannot fill the space occupied
by the ice which by its fusion has given birth to them; hence the
production of a vacuum, the inseparable companion of each liquid
flower.
The fragment of compact ice whose elements assume such
beautiful crystalline forms is itself a crystal. This was shown by Sir
David Brewster, who employed for the purpose of analysis that
modified form of light which we call polarised light. It is singularly well
adapted to bring out the peculiarities of the main structure of
substances, owing to the coloured figures which it outlines on a
screen after passing through them. All crystals with an axis—such,
for instance, as Iceland spar—yield a series of brilliantly-tinted rings,
traversed by a regularly-formed cross entirely black. As ice produces
the same figures, we are justified in attributing to it the same kind of
crystallization. We must note, however, that we are referring now to
the thick ice formed on our canals and lakes. If we examined the first
film formed on the surface of the water, we should discover in it a
completely irregular crystallization, the ray of polarised light
producing only a mosaic of varied tints, distributed without any order.
But it is easy to explain the way in which this primary crust or film is
produced. Those portions of the fluid mass in contact with the air are
the first to freeze, but each molecule of ice abandons its heat to the
contiguous water, which thereby is slightly raised in temperature, and
the result is a partial congelation. The surface we are examining then
presents a network of fine needles intercrossed in every direction,
and forming a kind of delicate lace, the meshes or intervals of which
are gradually filled up. When the network is transformed into a
continuous sheet, the loss of heat is diminished more and more as
this external crust grows thicker and thicker; but the development of
the ice invariably takes place by means of long interlaced needles,
as the reader may see for himself by breaking off a portion from the
nearest pond (in winter), and examining the sectional surface.
Having said thus much in reference to the crystallization of ice
and snow, we proceed to explain the regelation and moulding of ice.
Some years ago, Faraday astonished the scientific world by a very
curious experiment. Splitting into two parts a piece of ice, he brought
together the parts at the moment that fusion took place on their
surfaces, and they united immediately. How are we to account for
this effect, which can be produced even in hot water?
When the temperature of water rises, the surface molecules first
become liquid, then gaseous; being placed beyond the coercitive
action of the surrounding particles, they are easily set free;
transported, on the contrary, into the centre of the mass, they are
brought absolutely under the influence of this action, which induces a
new solidification,—or, to use the scientific term, a regelation. In this
way it becomes easy to understand how very various forms can be
communicated by simple pressure to a fragment of ice. If the
observer successively places a straight bar in moulds of increasing
curvature, he may easily compel it to assume the shape of a ring or
even of a knot. In each mould, it is true, the ice breaks; but if the
pressure is kept up, the surfaces of the fragments are brought into
contact, and adhere so as to re-establish a condition of continuity. A
snowball may thus be converted into a sphere of ice, and the sphere,
by constant pressure, into a cup or a statue.
Professor Tyndall refers to a remarkable instance of regelation
which he observed one day in early spring. A layer of snow, not quite
two inches thick, had fallen on the glass roof of a small conservatory,
and the internal air, warming the panes, had melted the snow so far
as it was in immediate contact with them. The entire layer had
slipped down the pane, and projected beyond the edge of the roof,
without falling, and had bent and curved as required, just like a
flexible body.
MOULDING ICE.

The snow-fields which overspread the upper part of every glacier,


whether in the Arctic Regions or elsewhere, are composed of
crystallized snow, whose fragile, delicate, and fairy-like architecture
endures so long as it remains dry, but undergoes a great
transformation when the sun, melting the upper stratum, allows the
water to interpenetrate its substance. The fluid, congealing anew
during the night, transforms the snow into the condition technically
known as névé; a term given by the Swiss physicists to a granular
mass composed of small rounded icicles, disaggregated, but more
adhesive than snow-flakes, and of a density intermediate between
that of snow and that of ice. Under the pressure of new layers, and
as a result of infiltrations of water, the névé unites, and solders into
ice of constantly increasing compactness.
But glacier-ice presents some other curious peculiarities. Every
abundant snow-fall on the summit of the mountains forms a layer
easily distinguishable from preceding layers—which, in most cases,
have already passed into the névé condition. This stratification
becomes more apparent when the whiteness of the surface has
been sullied by dirt or dust wafted on “the wings of the wind.” It is
perceptible also in ice; but here we must not confound it with another
phenomenon of which the cause is different, the veined structure.
In places where glaciers have been accidentally cut down in an
almost vertical direction, the section is found to exhibit a series of
parallel veins, formed by a beautiful and very transparent azure ice
in the midst of the general mass, which is of a whitish colour, and
slightly opaque.
In different glaciers, and in different parts of the same glacier,
these blue veins will vary in number and intensity of colouring. They
are specially beautiful in crevasses of recent formation, and on the
sides of channels excavated in the ice by tiny rills resulting from
superficial fusion. Not a few glaciers exhibit this remarkable veined
structure throughout their entire extent. When a vertical cutting
exposes the delicate azure network to atmospheric influences, the
softer ice melts prior to the fusion of the blue ice which then remains
in their detached leaflets. On examining these attentively, we cannot
fail to remark the absence, or, at all events, the extreme rarity, of air-
bubbles, though they are so plentiful in the coarser ice.
Professor Tyndall’s explanation of this phenomenon is as
interesting as it is ingenious. While on a visit of inspection to the
slate-quarries of Wales, he had occasion to study the cleavage of
the rocks which compose them; in other words, their faculty of
dividing naturally, a property inherent in all crystals. The schistous
slate separates easily into sheets, and in traversing different quarries
one sees that all the planes of cleavage are parallel in each. From
this circumstance our men of science were at first induced to look
upon slates as the products of the stratification of different deposits.
Such an explanation, however, could not be accepted by Tyndall,
when he observed that the minute fossils embedded in them were
constantly misshapen and flattened in the direction of the plane of
cleavage, because the great modification they had undergone could
not have taken place in superimposed strata at the bottom of the
primeval sea. He concluded that these schists, therefore, must have
been subjected to a considerable pressure; and further, that this
pressure must have been exercised at right angles with the plane of
separation of the different layers.
A long series of experiments proved that many bodies, when
forcibly compressed, exhibit in their structure a very distinctly marked
lamination, and frequently veins of very great beauty.
He carefully examined iron which had passed under the steam-
hammer, or through the rolling-mill; clay and wax were subjected to
the hydraulic press. In all cases he detected signs of cleavage; and
hence we are justified in the inference that the phenomenon is
invariably produced by pressure in all bodies of irregular internal
structure. Such is the result with glacier-ice, from whose mass the
air-bubbles introduced by the snow are gradually expelled. At first of
brilliant whiteness, it assumes, in the parallel layers corresponding to
the planes of cleavage, those beautiful azure tints which characterize
the veined structure. So little has it to do with stratification, that in
places where this is apparent it has given rise to a series of
horizontal lines, while the parallel veinings, in the same masses of
ice, are all inclined at an angle of about 60°.
The tendency to cleavage in compact ice would seem to explain
the regular form of those fragments or detached pieces with which
some parts of the glaciers are covered. Usually they occur as cubes,
or as rectangular parallelopipeds. The Alpine mountaineers name
them séracs,—in allusion to their resemblance to certain cheeses
which bear this name, and which are manufactured in rectangular
boxes. They have been found in many parts of a really colossal size,
measuring fifty feet in length, breadth, and depth, and as regular in
shape as if they had been hewn with a chisel.
There are many interesting points connected with the formation
and constitution of glaciers which we should gladly discuss, but we
are confined by our limits to remarks of a general character, and we
must now pass on to speak of the phenomena attendant upon their
motion. No doubt, the traveller who for the first time comes in sight of
one of these huge ice-rivers, and sees the mighty mass apparently
rooted to its valley-bed, solid, unchangeable, adamantine, finds it
hard to believe that it moves onward with a certain and an unresting,
though a gradual progress. It looks like a noble river, suddenly
petrified by some overwhelming force: congealed, as it flowed, in a
moment, by some irresistible spell! Such, indeed, is the conception
of the poet:—
“Ye ice-falls! ye that from the mountain’s brow
Adown enormous ravines slope amain....
Torrents, methinks, that heard a mighty voice,
And stopped at once amid their maddest plunge!
Motionless torrents! silent cataracts!”

And this conception is justified by the aspect of the glacier. Thus,


of the Glacier du Géant, Professor Tyndall says:—“It stretches
smoothly for a long distance, then becomes disturbed, and then
changes to a great frozen cascade, down which the ice appears to
tumble in wild confusion. Above the cascade you see an expanse of
shining snow, occupying an area of some square miles.” But we shall
see that here, as in the world of man, appearances are deceitful, and
that the glacier well deserves to be called an ice-river, in allusion to
its regular and continuous motion.
Between the snow-fall in the higher regions of the globe, and the
quantity of snow which every summer disappears through
liquefaction, the difference is very considerable. The supply, so to
speak, exceeds the demand, and a residuum is annually left. It is
only below the perpetual snow-line that the snow created and
accumulated in winter is wholly melted in the warm season. And,
therefore, if for any considerable period the excess upon any
particular mountain continued to accumulate, immense masses of
ice would gradually rise to the extreme height in the atmosphere
affected by aqueous phenomena.
Rendu, the Roman Catholic prelate, who first led the way to the
discovery of the true nature of glaciers, says, very justly,—“The
economy of the world would be soon destroyed, if at certain points
accumulations of matter prevailed. The centre of gravity of the globe
would be insensibly displaced, and the admirable regularity of its
movements would be succeeded by disorder and perturbation. If the
Poles did not send back to the Equatorial seas the waters which,
reduced into vapour, issue daily from these burning regions, to be
converted into ice in the Arctic and Antarctic Zones, ocean would be
drained dry, and life would cease, as well as water, to circulate
throughout our world. The Creator, however, in order to ensure the
permanence of His almighty work, has called into existence the vast
and powerful law of circulation, and this law the careful observer
sees reproduced in all the economy of Nature. The water circulates
from the ocean into the air, from the air it spreads over the earth, and
from the earth it passes into the seas. The rivers return from whence
they came, in order that they may issue forth anew; the air circulates
around the globe, and, as it were, upon itself, passing and repassing
successively at all the altitudes of the atmospheric column. The
elements of every organic substance circulates in changing from the
solid to the liquid or aëriform state, and in returning from the latter to
the state of solidity or organization. It is not improbable that the
universal agent which we designate under the name of fire, light,
electricity, and magnetism, has probably also a circle of circulation
as extensive as the universe. Should its movements ever be known
to us more than they now are, it is probable that they would afford
the solution of a host of problems which still defy the intellect of man.
Circulation is the law of life, the method of action employed by
Providence in the administration of the universe. In the insect, as in
the plant, as in the human body, we find a circulation, or rather
several circulations,—blood, humours, elements, fire, all which enter
into the composition of the individual.”
However fanciful may be some of the amiable prelate’s
speculations, it is certain that the glaciers obey this law of circulation.
The snow-accumulations in the upper regions are to some extent
reduced by the descent of the avalanches,—that is, of masses of
snow and ice which detach themselves from the mountain-sides and
dash headlong into the valleys below, where they are rapidly melted
by the warmer atmosphere. But this would, in itself, be wholly
insufficient. Another movement, at once more efficacious and more
regular, is necessary; a movement which embraces the entire
system of the ice-masses, and which carries the glaciers below the
perpetual snow-line, so that every year they may give up a portion of
their terminal extremities. The discovery of this general progression
is one of the most fertile with which, of late years, the physics of the
globe have been enriched.
Professor Tyndall rightly observes that there are numerous
obvious indications of the existence of glacier-motion, though it is too
slow to catch the eye at once. The crevasses change within certain
limits from year to year, and sometimes from month to month; and
this could not be if the ice did not move. Rocks and stones also are
observed, which have been plainly torn from the mountain-sides.
Blocks seen to fall from particular points are afterwards noticed lower
down. On the moraines rocks are found of a totally different
mineralogical character from those composing the mountains right
and left; and in all such cases strata of the same character are found
bordering the glacier higher up. Hence the conclusion that the
foreign boulders have been floated down by the ice. Further, the
ends or “snouts” of many glaciers act like ploughshares on the land
in front of them, overturning with irresistible energy the huts and
châlets that lie in their path. Facts like these have been long known
to the inhabitants of the High Alps, who were thus made acquainted
in a vague and general way with the motion of the glaciers. But
Science cannot deal with generalities: it requires precise and
accurate information; and this information, so far as the progression
of the glaciers is concerned, has been obtained through the patient
labours of Rendu, Charpentier, Agassiz, Desor, Vogt, Professor
Forbes, Bravais, Charles Martins, Hopkins, Professor Tyndall,
Colomb, John Ball, and Schlagintweit. Their experiments and
observations have established the truth of certain immutable
principles, and proved the existence of a general law of movement.
The accumulation of the débris hurled headlong by the mountains
forms on the glacier-surface long lines of stone and earth, which are
called moraines; these diverge in certain directions, according to the
circumstances we now come to explain.
The landslips which occur on the banks or edges of the glacier
give rise to the lateral moraines, which are enlarged and extended
daily by the twofold effect of the fall of stones and débris, and the
progressive movement which carries them along with the whole
mass of ice. Towards the centre of the great glaciers, in almost every
case, is found a medial moraine; the result of the encounter of the
lateral moraines of two glaciers which have united into one. These
superficial moraines participating in the movement of the glacier,
each of their blocks eventually rolls to the foot of the terminal
precipice, and thus a frontal moraine is formed on the very soil of the
valley, like an embankment raised to prohibit the further advance of
the ice. And, lastly, the bed of sand, gravel, pebbles, and detritus
which is found beneath the glacier, and over which it glides, is called
the profound moraine.
The furrows wrought by this last-named stratum on the bottom of
the glacier-channels show the wonderful force of friction which the
glacier exercises during its descent. The depths of these furrows
depends entirely on the hardness of the débris carried down by the
glacier, and the nature of the rocks submitted to the friction. The
polish assumed by these rocks when they are sufficiently solid to
resist the thunderous march of the glacier, indicates the enormous
pressure which it exercises on the slopes of the valley through which
it forces its way. This effort, bearing principally on the side of the
rocks turned in the direction of their crests, impresses upon them a
peculiar rounded form, so like the appearance of a flock of sheep
(moutons) that De Saussure gave them the name of roches
moutonnées.
Connected with the scientific evidence of the progressive
movement of glaciers, a glacier in the Bernese Oberland will for ever
be memorable. Two branch glaciers, the Lauteraar and the
Finsteraar, unite at a promontory called the Abschwung to form the
trunk-glacier of the Unteraar, which carries a great medial moraine
along its colossal back.
Here in 1827, an “intrepid and enthusiastic” Swiss professor,
Hugi, of Solothurm (or Soleure), erected a small cabin of stones for
the purpose of observations upon the glacier. The hut moved, and he
took steps to measure its motion. In three years, 1827 to 1830, it
moved 330 feet downwards. In 1836 it had descended 2354 feet;
and in 1841, it had accomplished a journey of 4712 feet. [This was at
the rate of about 336 feet a year.]
In 1840, M. Agassiz, with some scientific friends, Messrs. Desor,
Vogt, and Nicolieb, established themselves under a great
overhanging slab of rock on the same moraine, and by means of
side walls, and other appliances, constructed a rough abode which,
because some of these men of science came from Neufchâtel, they
named the “Hôtel des Neuchâtelois.”
In two years after its erection, Agassiz discovered that it had
moved downwards no less a distance than 486 feet.
These and some similar measurements brought to light a very
important fact. The reader will observe that the middle numbers,
corresponding to the central portion of the glacier, are the largest:
hence it was obvious that the centre of a glacier, like that of a river,
moves more rapidly than the sides.
Owing to the greater central motion of a glacier, its crevasses
invariably assume a curved outline, of which the convexity advances
towards the bottom of the valley.
It has also been ascertained that the superficial part of a glacier
moves more rapidly than its base.
Again: Tyndall and Hirst, by employing instruments of great
precision, have demonstrated that the maximum of motion is not to
be found exactly in the centre, but that, according to the windings of
the valley through which the glacier flows, it moves sometimes to the
right of the centre, and sometimes to the left. Now, the progression
of a river exhibits all the characters we have just enumerated, and
the truth foreshadowed by Rendu has been confirmed in every
detail. The glacier is a “river of ice.”
The reader will naturally ask, How can a substance of such
apparent rigidity as ice obey, as it does obey, the same laws which
regulate the movement of fluids? I can understand, he may say, how
water flows in such and such a manner: it is a liquid, and its
molecules are deficient in the property of cohesion; but that so solid,
and firm, and unimpressible a substance as ice should be capable of
motion seems impossible. I can understand very easily that a mass
of ice, when loosened or detached from its resting-place, will glide
downwards until arrested by some adequate obstacle; but this is not
the kind of motion you are describing. According to your
explanations, every constituent portion of the glacier moves, and the
central faster than the lateral, and the surface faster than the base.
These objections were advanced by men of science when the
motion of glaciers was first put forward as a theory; and the answer
given by Scheuchzer was, that a glacier might be compared, in the
summer season, to a sponge saturated with water, which, when
afterwards congealed by the cold temperature of autumn and winter,
expanded, and produced a dilatation of the mass in every direction.
Then, as it could not recede, as it could not reascend its valley-
slope, the augmentation of size would necessarily take place in its
lower portion.
It is unnecessary for us to explain why this answer was
unsatisfactory. Subsequent observations, however, proved its
impossibility, and Professor Forbes then put forward his ideas of the
viscous character of ice. But these, too, did not meet the conditions
of the phenomenon; and the view now adopted is that of Professor
Tyndall, who has shown that it is the result of the regelation we have
already described.
Professor Forbes enunciated his theory in words to the following
effect: “A glacier is an imperfect fluid or viscous body, which is urged
down slopes of certain inclination by the natural pressure of its
parts.” But we know the exceeding brittleness of ice, and how is
viscosity compatible with brittleness? We know, too, that crevasses
and fissures will suddenly form on a glacier, like the cracks on a
pane of glass. But if ice were viscous, and could expand, dilate, or
stretch as viscous substances do, these crevasses would be
impossible. They would gradually close up, like an indent in a mass
of jelly. And yet it cannot be denied that a glacier does move like a
viscous body; the centre flowing past the sides, the top flowing over
the bottom, while the motion through a curved valley corresponds to
fluid motion. How are we to reconcile these apparently conflicting
circumstances?
By Professor Tyndall’s regelation theory, which is founded on a
fact already mentioned; namely, that when two pieces of thawing ice
are brought in contact, they freeze together.
This fact, and its application irrespective of the cause of
regelation, may be thus illustrated: “Saw two slabs from a block of
ice, and bring their flat surfaces into contact; they immediately freeze
together. Two plates of ice, laid one upon the other, with flannel
round them overnight, are sometimes so firmly frozen in the morning
that they will rather break elsewhere than along their surface of
junction. If you enter one of the dripping ice-caves of Switzerland,
you have only to press for a moment a slab of ice against the roof of
the cave to cause it to freeze there and stick to the roof.
“Place a number of fragments of ice in a basin of water, and
cause them to touch each other; they freeze together where they
touch. You can form a chain of such fragments; and then, by taking
hold of one end of the chain, you can draw the whole series after it.
Chains of icebergs are sometimes formed in this way in the Arctic
seas.”
From these observations we deduce the following result:—Snow
consists of small particles of ice. Now, if by pressure we squeeze out
the air entangled in thawing snow, and bring the little ice-granules
into close contact, they may be expected, as they do, to freeze
together; and should the expulsion of the air be complete, the
squeezed snow will assume the appearance of compact ice.
It is in this way that the consolidation of the snows takes place in
the Arctic as in the higher Alpine regions. The deeper layers of the
névé are converted into more or less perfect ice by the pressure of
the superjacent layers; and further, they are made to assume the
shape of the valley which they fill, by the slow and continuous
pressure of its sides.
In glaciers, as Professor Tyndall points out, we have ample
illustrations of rude fracture and regelation; as, for example, in the
opening and closing of crevasses. The glacier is broken on the
cascades, and mended at their bases. When two branch glaciers lay
their sides together, the regelation is so firm that they begin
immediately to flow in the trunk glacier as in a single stream. The
medial moraine gives no indication by its slowness of motion that it is
derived from the sluggish ice of the sides of the branch glaciers.
We may sum up the regelation theory in few words. The ice of
glaciers changes its form and retains its continuity under pressure
which keeps its particles together. But when subjected to tension,
sooner than stretch, it breaks, and behaves no longer as a viscous
body.
These are Professor Tyndall’s words, and the fact which they
embody it would be difficult to set forth more clearly or more
concisely.

A POLAR GLACIER.
Having said thus much of the structure, causes, characteristics,
and movement of glaciers, we proceed to consider some of the more
remarkable of those which are situated in the Arctic World.
The glaciers of the Polar Regions do not differ in structure or
mode of formation from those of other countries. Yet they possess
some peculiar features, and to a superficial observer might seem
independent of the physical laws we have attempted to explain. That
this is not the case has been shown by Charles Martins, who
carefully studied the glaciers of Spitzbergen on the occasion of the
exploring voyage of the Recherche to that island, and has
demonstrated that their differences are but a particular case of the
general phenomenon.
As special characters he points out, first, the rarity of needles and
prisms of ice, which he attributes to the slight inclination and the
uniformity of the slopes, as well as to the diminution of the solar
heat, which, even in the long summer days, does not melt the
surface. There are no rills or streams capable of hollowing out
crevasses and moulding protuberances or projections. But
transversal crevasses produced by the movement of the glaciers are
numerous, and these are often very wide and very deep.

GLACIER, ENGLISH BAY, SPITZBERGEN.


In the terminal escarpment, which melts in proportion as it
plunges into the sea, immense caverns are sometimes seen;
caverns so immense that the azure-gleaming grottoes of the
Arveiron and Grindelwald, so much admired by European travellers,
are but miniatures. “One day,” says Charles Martins, “after having
ascertained the temperature of the sea off the great glacier of Bell
Sound, I proposed to the sailors who accompanied me to carry our
boat into its cavern. I explained to them the risk we should incur,
being unwilling to attempt anything without their consent. When our
boat had crossed the threshold, we found ourselves in an immense
Gothic cathedral; long conical-pointed cylinders of ice descended
from the roof; the recesses seemed so many chapels opening out of
the principal nave; broad fissures divided the walls, and the open
intervals, like arches, sprang towards the summits; azure gleams
played over the icy surface, and were reflected in the water. The
sailors, like myself, were dumb with admiration. But a too prolonged
contemplation would have been dangerous; we soon regained the
narrow opening through which we had penetrated into this winter
temple, and, returning on board our vessel, preserved a discreet
silence respecting an escapade which might have been justly
blamed. In the evening, we saw from the shore our cathedral of the
morning slowly bend forwards, detach itself from the parent glacier,
crash into the waves, and reappear in a thousand blocks and
fragments of ice, which the retiring tide carried slowly out to sea.”
The Spitzbergen glaciers do not exhibit those numerous
moraines which are observed on the majority of those of
Switzerland.
The mountains, not being very lofty, are buried, as it were, under
their burden of glaciers, instead of preponderating over them, and
seem with difficulty to lift their peaks out of the mass of ice and snow
surrounding them. Consequently, there are no considerable landslips
or falls of earth and stone, which, accumulating along the borders of
the glaciers, might form moraines. Martins is of opinion that the
Spitzbergen glaciers correspond to the upper part of the glaciers of
Switzerland; to so much, that is to say, as lies above the perpetual
snow-line.
Now, he says, the higher we ascend on an Alpine glacier, the
more do the lateral and medial moraines diminish in width and form,
until they taper away and finally disappear under the high névés of
the amphitheatres from which the glacier issues, just as the
mountain torrents often take their rise in one or in several lakes
terraced one above the other.

GLACIER, BELL SOUND, SPITZBERGEN.


For all these reasons, he adds, the medial and lateral moraines
are scarcely conspicuous on the glaciers of Spitzbergen; a number
of stones and boulders may be seen along their sides, and
sometimes in their centre, but the ice is never hidden, as in the Alps,
under the mass of débris accumulated upon it. As for the terminal
moraines, they must be sought at the bottom of the sea, since the
terminal escarpment nearly always overhangs it. Hence, the blocks
of stone fall simultaneously with the blocks of ice, and form a
submarine frontal moraine, of which the two extremities are
occasionally visible upon the shore.
In a previous chapter we have alluded to the manner in which
icebergs are formed by the detachment from the seaward extremity
of the glacier of huge masses of ice, which the current carries out
into the open sea. To the description already given, we may here add
that which Charles Martins furnishes in his valuable and interesting
record of persevering scientific enterprise, “Du Spitzberg au Sahara”:
—In Spitzbergen, he says, the glacier, after a traject of more or less
considerable duration, reaches the sea. If the shore be rectilineal, it
advances no further; but, in the recess of a bay, where the shore is
curved, it continues its progression, supporting its bulk on the sides
of the bay, and advancing above the water, which it overhangs. This
is easily understood. In summer the sea-water at the bottom of the
bays is always at a temperature a little above 32°; on coming in
contact with this comparatively warm water the glacier melts, and, at
low tide, an interval is perceptible between the ice and the surface of
the water. The glacier being no longer supported, partially crumbles
and gives way; immense blocks detach themselves, fall into the sea,
disappear beneath the water, reappear revolving on their own axes,
and oscillate for a few moments until they have taken up their
position of equilibrium. The blocks thus detached from the floating
masses, of all sizes and shapes, are called icebergs.

STEAMER “CHARGING” AN ICEBERG, UPERNAVIK, GREENLAND.


Our traveller records that twice a day, in Magdalena Bay and Bell
Sound, he was an eye-witness of this partial ruin of the extremity of
the glaciers. Their fall was accompanied by a noise like that of
thunder; the swollen sea rushed upon the shore in a succession of
gigantic waves; the gulf was covered with icebergs, which, caught in
the swirl and eddy, issued out of the bay, like immense fleets, to gain
the sea beyond, or were stranded here and there at points where the
water was shallow. The icebergs seen by M. Martins were not,
however, of any surprising magnitude; he estimates their average
height at thirteen to sixteen feet. We have seen that those of Baffin
Bay are tenfold more considerable and imposing; but then, in that
bay the temperature of the sea is below 32°; the glacier does not
melt when it enters the water; it sinks to the bottom of the sea; and
the portions detached from it are all of greater height than even the
submerged part of the icebergs which drift to and fro in the bays and
gulfs of Spitzbergen.

We may follow up this description with some observations by


Lieutenant Bellot, the chivalrous young Frenchman who perished in
one of the expeditions despatched in search of Sir John Franklin and
his companions. He is speaking of the masses of ice his ship
encountered soon after doubling Cape Farewell, the south point of
Greenland, and he remarks, that as Baffin Bay narrows towards the
south, the icebergs, first set in motion higher up the bay by the
northern gales, necessarily tend to accumulate in the gorge thus
formed, and so to impede and block up Davis Strait, even when the
higher waters are quite free. It is only through a series of alternate
movements of advance and recession that the bergs finally pass
beyond the barrier, and float out into the Atlantic, to undergo a slow
process of dissolution.
The mobility of the bergs, though necessary to navigation, forms
at the same time its peculiar danger, since a vessel is often placed
between the shore and the colossal masses driven forward by the
wind, or between these and the solid ice which as yet has not broken
up. It is useless to dwell upon the immense force possessed by
masses which are frequently several square leagues in extent, and
which, once in movement, cannot be stayed by any human
resistance. A sailing-vessel finds herself placed in conditions all the
more unfavourable, because the winds blow from the very direction
which she is bound to take in order to open up a way through the
floes. Now, if the gale is violent, it is perilous indeed to push forward
in the midst of a labyrinth of bergs, which form so many floating
rocks; if a calm prevails, a ship can move forward only by laborious
hauling or towed by the boats. The application of the screw-propeller
to steam-ships has given to them a great superiority, because they
are not liable to any accident to paddle-wheels, exposed as such
must be to collision with the floating ice. It is recorded that, on one
occasion, a screw-steamer, near Upernavik, on the coast of
Greenland, actually charged an iceberg, and drove right through it,
as a railway-engine might crash through a fence or hurdle. Of
course, the berg was of no great elevation; but its solid mass yielded
to the immense force of the steamship, and split into large
fragments.
In the convulsions caused by furious tempests, which are far from
being so rare within the Arctic Circle as is popularly supposed, the
shape of the bergs becomes very irregular, and the configuration of
the ice-fields is constantly undergoing modification. Hence it often
happens that the voyager sees before him an open basin of water of
greater or less extent, from which he is separated only by a narrow
strip of ice. In such a case he endeavours to effect an opening,
either by driving his ship at full speed against the weakest part of the
ice, or with the help of immense saws, twenty feet in length, which
are worked with a rope and pulley placed at the top of a triangle
formed of long poles; or, finally, by exploding a mine. When the ice is
not very solid, the ship is forced into the opening, against the sides
of which it acts like a wedge. It will sometimes occur, in the course of
the operation, that the ice-fields, set in motion by the wind or the
currents, close in together, after having treacherously separated for a
moment, and the vessel is then subjected to a dangerous pressure.
Unhappy the mariner who does not foresee or sufficiently note the
warning signs of this accident, which is almost always accompanied
by fatal consequences. The ice, which nothing can check, passing
underneath the ship, capsizes it,—or, if it resists, crushes it.

We have alluded to the colossal bergs of Baffin Bay. These are


thrown off from the northern glaciers, and particularly from the
enormous ice-river named after Humboldt, which cumbers the
declivities of the Greenland Alps, beyond the 79th parallel. It has
been a frequent source of surprise to navigators that these mighty
masses should float in a contrary direction to that of the ice-fields
which descend with the Polar current towards the Atlantic. They
reascend with such rapidity that they shatter the so-called “ice-foot,”
or belt of ice, still adhering to the shore. Captain Maury has collected
numerous observations on this important subject, and he quotes the
case of a ship which was being laboriously hauled against the
current, when an enormous floating mountain coming up from the
south steered against it, but fortunately did not come into collision
with it, and forging ahead, very quickly disappeared. How is such an
incident to be explained? By the existence of a submarine counter-
current, acting on the lower extremity of the submerged portion of
the berg, which, as we have stated, is always seven or eight times
larger than the bulk above the surface of the waves.
Our whalers, in their hazardous expeditions, often derive
assistance from these moving islands. They seek shelter under their
lee when sudden storms arise; for the huge bergs are scarcely
affected by the most violent gales. They find their shelter valuable
also during certain operations of the fishery for which rest and quiet
are necessary. Yet it is not absolutely exempt from danger. The
seeming friend may prove to be a concealed foe. The iceberg may
collapse, or be capsized; or formidable fragments, loosened from
their sides or summits, may topple headlong and threaten to
overwhelm the ship beneath: but as on these and other accidents we
have already dwelt at length, we refrain from wearying our readers
with a twice-told tale. The repetition in which, to some extent, we
have indulged, was needful, in order to show the reader in what way

You might also like