On the theory of advective effects on

biological dynamics in the sea



Proceedings of the Royal Society A, 453, 2295-2324, 1997.





Allan R. Robinson

Division of Engineering and Applied Sciences, Department of Earth and Planetary Sciences,

Harvard University, Cambridge, MA 01238, U.S.A.

robinson@pacific.harvard.edu



Abstract

A general theoretical approach to the study of advective effects on biological oceanographic dynamics is formulated and some simple examples are presented. The theory of characteristics provides a model which rigorously combines aspects of local biological dynamics with consideration of fluid particle flow along Lagrangian trajectories. A local kinematical analysis is carried out for gradients both of velocity and of biological state variables. This is followed by a study of the effects of a stretching deformation, which contains a frontogenetic effect, on one to three component biological systems. Problems include the approach to local biological equilibria and the stabilities of the equilibrium states; the competition between biological and advective equilibria; and nonlinear grazing processes and nutrient utilization.

1. Introduction

Biological dynamics in the sea involves interactions among physical, biological and chemical processes and coupled processes. A multitude of scales and phenomena are involved, many of which are as yet unknown. Research in interdisciplinary theory and modeling is accelerating, and relevant studies range from idealized isolated process studies to attempts at simulating complex and realistic oceanic dynamics. Concomitantly biological dynamical studies are carried out in zero, one, two and three spatial dimensions. One aspect of the interpretation of some zero-dimensional studies involves the concept of local biological dynamics and equilibria. In such cases the physics may be assumed to be unimportant or may be parameterized by scale-separation hypotheses.

Physical processes which influence the biology consists of physiological effects such as temperature-dependent processes, diffusion and mixing, and advection. Physical processes may compete with or enhance biological processes. Advective effects, which are perhaps least well understood theoretically, are isolated as the focus of the present study. Advection involves the movement of biological material via the physical flow field. Vertical and horizontal velocities transport and entrap dissolved and particulate matter over multiple scales in space and time. A large variety of relevant phenomenological flow structures occur in the ocean, and their associated advective effects may be best treated deterministically or statistically.

The present study adopts an idealized general theoretical approach to the study of advective effects. The attempt is to identify and expose some basic processes which may aid in the interpretation of simulations and which may form the basis of further more realistic studies. The focus is primarily upon the influence of a deformation gradient in the flow field on a dynamically active tracer field, and the results may have some general applications to e.g. chemical and engineering dynamics. The general model, which can be related to zero-dimensional biological dynamics and fluid particle trajectories, is introduced in §2. Sections 3, 4 and 5 subsequently deal, respectively, with advective, biological and interactive processes. Section 6 summarizes and concludes.



2. The model

(a) Formulation

Consider an n-component biological system with the state variables

fi(x,y,z,t),          i = 1,¼, n     (2.1)
for the mass densities of plankton, nutrients, etc. in suitable units of mass (m) divided by volume (l3). The state variables are assumed to be continuous fields (Batchelor, 1967, Chapter 1), diffusion is neglected, and the necessity of Reynolds-averaging of physical and biological turbulent fluctuations in the fluid (Hinze, 1975; Okubo, 1980, Chapter 10) is ignored. If, in the absence of motion, the biology is governed by the zero (spatial) dimensional dynamics

dfi
dt
= Bi(f1,¼,fn)       i = 1,¼,n
º Bi(fj)                   j = 1,¼,n,
(2.2)
then the dynamics, in the presence of a flow field, is governed by
Dfi
Dt
= B(fj).     (2.3)
Note that the argument fj of B symbolically represents the arbitrary individual dependencies of B upon f1¼fn.

The advective derivative in Cartesian coordinates is

Dfi
Dt
= fi
t
+u fi
x
+v fi
y
+w fi
z
(2.4a)
= fi
t
+
x
(ufi)+
y
(vfi)+
z
(wfi),
(2.4b)
where the equivalence of the two forms of eq. (2.4) arises from the conservation of mass, or continuity equation, of an incompressible liquid
u
x
+ v
y
+ w
z
= 0.     (2.5)
In most cases of interest we will represent biological activity by simple self-interactions and predator-prey interactions (Murray 1993, Chapter 3), with

Bi(fj) = fi(ai0+ n
å
j = 1 
aijfj  ).     (2.6)

The advective derivative D/Dt, as given by equation (2.4a), represents the time rate of change of fi following the fluid motion, but expressed in the Eulerian viewpoint for a space-time field (equation (2.1)) at a fixed point in the flow (Lamb, 1932, Chapter 1; Batchelor, 1967, Chapter 2). From the Lagrangian viewpoint, i.e., following tagged fluid particles along finite trajectories, zero-dimensional biological dynamics is simply occurring within each particle. Equation (2.4b) presents an alternative useful interpretation of the advective process at a fixed Eulerian point from a flux viewpoint. Taking into account all processes which act as effective sources (or sinks), equations (2.2) and (2.3) are simply point-wise statements of the conservation of fi. The local advective flux (units of mass per unit area per unit time -ml-2t-1) is

FA = ufi ^
i
 
++vfi ^
j
 
+wfi ^
k
 
     (2.7)
and its divergence (convergence) Ñ· FA can provide a local sink (source) of biomass. The flux divergence process is illustrated for an infinitesimal Cartesian volume in figure 1. However, recall that the divergence theorem in general relates the net normal flux through the surface S of an arbitrary volume V to the volume integral of the divergence, and that for an infinitesimal volume
Ñ· F =
lim
V® 0 
1
V
ó
õ


S 
^
n
 
· F  dS     (2.8)
is the mass outflow per unit volume per unit time.




Figure 1

Figure 1. At the fixed point (x,y,z) the net difference of rate of inflow and outflow of f through the six faces tends to change the amount of f within the infinitesimal volume.




In the present study we will restrict ourselves to two (spatial) dimensional flows, whence continuity implies the existence of a stream function y(x,y,t) such that

u(x,y,t) = - y
y
  ,       v(x,y,t) = y
x
     (2.9)
satisfy equation (2.5) identically. The x-y plane can be regarded as a horizontal surface or a vertical section in the ocean. In the former case, if the flow is of large enough scale to be in geostrophic balance (Pond and Pickard, 1978), the stream function is related to the geostrophic pressure (p) via p = r0fy, where r0 is the mean density and f = 2Wsinq is the Coriolis parameter (W is the earth's rotation and q is the latitude).

It will be convenient to express the model in both dimensional (2.10a) and nondimensional (2.10b) forms. For the latter case: the fi will all be scaled by a representative mass density M; time by t0; distance by L; and speed by u0. Since the context will be clear, we will use identical symbols for both the dimensional and nondimensional versions for the dependent and independent variables. Thus the final model takes the forms

fi
t
+u fi
x
+v fi
y
= Bi(fj) = fi(aio+ n
å
j = 1 
aijfj)
(2.10a)
fi
t
+a[u fi
x
+v fi
y
] = Bi(fj) = fi[aio+ n
å
j = 1 
aijfj]
(2.10b)
In eq. (2.10b), the nondimensional parameters

a º u0t0
L
  ,      aio º t0aio  ,      aij º Mt0aij     (2.10c)
characterize the relative importance of local time changes (f/t) due respectively to advection, net birth or death, self-interaction (aii), and predator-prey interactions (grazing).

The left-hand side alone represents local time change plus the advection of the scalar field and the right-hand side alone, the ``zero-dimensional'' local biological dynamics. In general, scalar fields are characterized as passive or active depending upon whether or not they influence the flow field, and inert or dynamical depending upon whether they are simply advected or have sources and sinks due, e.g., to radioactivity, chemical reactions or biological activity.

A version of the parameter a (denoted S) based upon the growth rate t0 = ai0-1 was first identified by Obrien and Wroblewski (1973) as an appropriate measure of advective effects. Platt and Denman (1975) and Platt et al. (1977) scaled a general biomass equation and estimated ranges of relative ratios from data available at that time, which are generally still relevant. A useful quantitative overview of the nondimensional approach is also presented by Wroblewski (1983). A recent example of a nondimensional analysis of scale-dependent physical and biological processes governing the distribution and variability of biomass for a particular species is provided by Horne and Schneider (1994).

(b) Solution

Equations (2.10) are a set of n-coupled first-order partial differential equations. Since the left-hand sides are linear in the fi, and are identical operators on each of the fi, it is a quasilinear set with the same principal parts which can be solved symbolically via the method of characteristics (Courant and Hilbert, 1989; Chapter II, Appendix I). Consider a solution given in terms of a parametric variable s in the form

fi(x(s),y(s),t(s))  .     (2.11 a)
Then if

dfi
ds
= fi
x
dx
ds
+ fi
y
dy
ds
+ fi
t
dt
ds
  ,     (2.11b)
and if

dt
ds
= 1  ,       dx
ds
= u(x,y,t)  ,       dy
ds
= v(x,y,t)     (2.12a)
and also

d fi
ds
= Bi(f1¼fn)  ,     (2.12b)
then equation (2.11) solves equation (2.10). The solutions to (2.12) are called the family of characteristic curves belonging to equations (2.10)

The parametric initial value problem specifies data, without loss of generality, at s = 0. The initial values of all the variables are to be specified in terms of two additional parametric variables (r,q), since there are three independent variables in each of equations (2.10a,b). We designate

t(s = 0;r,q) = a(r,q)  ,      x(0,r,q) = b(r,q)  ,      y(0,r,q) = c(r,q)     (2.13a)
and

fi(0,r,q) = di(r,q)  .     (2.13b)
The formal procedure is to solve (2.12a) subject to conditions (2.13a) to yield

t(s,r,q),    x(s,r,q),    y(s,r,q)     (2.14a)
and then to invert to obtain

s(x,y,t),    r(x,y,t),   q(x,y,t)  .     (2.14b)
The set of equations (2.12b) is solved subject to conditions (2.13b) to yield
fi(s,r,q)     (2.15a)
which upon insertion of the parametric variables (2.14b) provides the biological fields in physical space

fi(x,y,t)  .     (2.15b)

The structure of the formal solution of the preceding paragraph is dynamically interesting and provides a valuable framework for interpreting coupled biological and physical processes. Advective processes appear in (2.12a) directly in terms of the Lagrangian trajectory equations but in the s-domain. Similarly, the zero-dimensional biological dynamics appear in the s-domain in eq. (2.12b). A special but important class of problems involves the specification of initial fields in the time domain, i.e., at t = 0. The first of (2.12a) is always directly integrable to t = s+a(r,q), and for this special case we have a = 0, s = t. Now (2.12b) is the local biological dynamics, but the local ``biological equilibrium'' (2.15a) can be altered significantly by the advective effects through r,q as given by (2.14b).

To illustrate the solution procedure, we consider the simplest example of the exponential (Malthusian) growth of phytoplankton (fi) with unlimited light and nutrient in the presence of a spatially uniform and constant velocity field (u0,vo). We seek the time evolution of an initial spatial distribution of f. Equations (2.12a,b) with reference to (2.10a) become

dt
ds
= 1       dx
ds
= u0       dy
ds
= v0       df1
ds
= a10f1     (2.16a)
with solutions (2.14a, 2.15a)

t = s+a,    x = u0s+b,   y = u0s+c   f1 = d1ea10s.     (2.16b)
The time domain initial condition
f1(x,y,t = 0) = d(x,y)     (2.16c)
requires for solution to equations (2.13)

a = 0,    b = r,    c = q     (2.16d)
whence (2.14b) becomes simply

s = t    r = x-u0t   q = y-v0t  .     (2.16e)
The final solution (2.15b)
f1 = d1(x-u0t,y-v0t)ea10t     (2.16f)
is an advective translation of the initial pattern with the plankton growing uniformly everywhere. The simple structure of (2.16f) as a product of advective effects times biological effects is generally deceptive as will be seen in Section 5.

An alternative formulation of the model in terms of the stream function defined in equation (3.2) can be useful. We consider a time-independent stream function y(x,y) as an independent variable and seek fi(y,y,t). Since

f
x
ê
ê
ê


y,t 
= f
y
ê
ê
ê


y,t 
· y
x
ê
ê
ê


y 
;    f
y
ê
ê
ê


x,t 
= f
y
ê
ê
ê


y,t 
· y
y
ê
ê
ê


x 
+ f
y
ê
ê
ê


y,t 
  ,     (2.17a)
equation (2.10a) becomes with (y,y) as independent variables

fi
t
+v(y,y) fi
y
= Bi(fj)  .     (2.17b)
Equation (2.17b) contains the familiar result that a time-independent distribution of a passive inert scalar can be any functional F(y) of the stream function field. More generally, if a passive dynamic scalar is distributed on the streamlines of the flow, it will evolve everywhere according to zero-dimensional dynamics, i.e. (2.17b) with [(fi)/(y)]|y,t = 0 reduces to equation (2.2). The alternative to equation (2.12) becomes

dt
ds
= 1       dy
ds
= v(y,y)       dfi
ds
= Bi  .     (2.17c)
Since (2.17b) now has only two explicit independent variables, only one additional parameter (r) is needed for the solution to (2.17c); y is a parameter of the solution. However initial x-dependence must be represented as x(y,y).



3. Advective processes

In this section we consider solutions to eqs. (2.10, 2.12) with Bi = 0 and (u,v) specified, i.e., the advection of a passive inert scalar. Even under these simplifications, analytical solutions for the trajectories are difficult and the flow of a tracer patch can be quite complex due to the stretching and folding of an initial distribution. Even smooth flows can give rise to chaotic advection (Ottino; 1989, 1990). Interesting examples of the complex trajectories that occur in oceanic mesoscale waves and coherent vortices are presented by Flierl (1981) and Dewar and Flierl (1985).

(a) Local analysis

Consider the initial value problem in the time domain. To provide some general insight we expand both an initial scalar tracer field and the advecting velocity in Taylor series in the neighborhood of the point (0,0). Since the tracer fields are inert and uncoupled we drop the subscript i. Thus

f(x,y,0) = c0+c1x+c2y     (3.1a)
u
= u0+u1x+u2y
(3.1b)
v
= v0+v1x+v2y
If the flow is nondivergent, then v2 = -u1 and a stream function exists. The case of pure translation is given by eq (2.16f) with a10 = 0. Removing the translation, the solution equations (2.14)-(2.15) to equations (2.11)-(2.12) are given by
f = c0+c1r+c2q     (3.1c)
s = t
r = é
ê
ë
coshgt- u1
g
sinhgt ù
ú
û
x- é
ê
ë
u2
g
sinhgt ù
ú
û
y     (3.1d)
q = - é
ê
ë
v1
g
sinhgt ù
ú
û
x+ é
ê
ë
coshgt+ u1
g
sinhgt ù
ú
û
y
where
g º [u12+u2v1]1/2  .
Note that at t = 0, r = x, q = y. Asymptotically for real g as t® ¥,
r = egt
2
é
ê
ë
æ
ç
è
1- u1
g
ö
÷
ø
x- u2
g
y ù
ú
û
     (3.1e)
q = egt
2
é
ê
ë
- v1x
g
+ æ
ç
è
1+ u1
g
ö
÷
ø
y ù
ú
û
  .
Thus a mechanism for the intensification with time of the initial tracer gradient is inherent in the shear of the flow field. However the general solution (3.1) is quite complicated, and g may be real or imaginary.

Insight into basic advective processes can be gained by a decomposition of the velocity gradient into its fundamental kinematic parts: divergence (x), stretching deformation (n), vorticity (w), shearing deformation (m)

x
= ux+vy = u1+v2 = 0
n
= ux-vy = u1-v2 = 2u1
w
= vx-uy = v1-u2
m
= vx+uy = v1+v2 = v1-u1
     (3.2)
as is traditional in fluid dynamics (Batchelor, 1967), meteorology (Stewart, 1945) and oceanography (Kirwan et al., 1994). The pure effects of each type of motion are shown on figure 2. The divergence and vorticity are invariant with respect to coordinate frame rotations, but not the deformations. However, a rotation through an appropriate angle into the principal axes of deformation will zero the shearing deformation. Thus here we need be concerned only with the effects of vorticity and stretching deformation.




Figure 2

Figure 2. Fundamental kinematic components of a velocity gradient: (a) divergence; (b) stretching deformation; (c) vorticity; (d) shearing deformation.




To illustrate the effects of vorticity we set u1 = 0 and u2 = -v1, whence w = 2v1. Then g is imaginary and

r
= xcosv1t+y sinv1t
q
= -xsinv1t+y cosv1t
     (3.3)
Inserting (3.3) into (3.1c) shows that the tracer gradients are simply rotated in space with period (p/w).

To illustrate the effects of deformation we set v1 = u2 = 0; and the solution is now

r = xe-u1t,      q = yeu1t,     (3.4a)
f = c0+c1xe-u1t+c2yeu1t.     (3.4b)
Referring to figure 2(b), it can be seen that the hyperbolic flow y = -u1xy is towards the x-axis (axis of contraction) and away from the y-axis (axis of dilation). The effect is to increase initial y-gradients of the tracer and to decrease initial x-gradients. As the flow continues in time, the gradient intensifies without limit.


(b) Frontogenesis

The role of deformation fields as a ``trigger'' mechanism for the generation of physical fronts in the atmosphere, or the signal that frontogenesis is occurring has been known for some time (Bergeron, 1928; Petterson, 1935). The analogous mechanisms is known to act as a causal mechanism for frontogenesis in the oceans (Simpson and James, 1986). Oceanic fronts have been identified as locations of intensified biological activity because of frontal upwelling of nutrients. It should also be recognized that the processes that cause strong physical gradients of e.g. temperature, will also tend directly to concentrate and cause strong gradients of biological and chemical tracers, i.e., to generate biological fronts. Convergence (regions of negative x) is also a frontogenetic kinematic mechanism, but the thin aspect ratios of the atmosphere and ocean (depth much less than horizontal scales) render them almost horizontally nondivergent and the process is less important.

Saucier (1953) describes a classical meteorological method for analyzing the tendency for frontogenesis from the horizontal deformation of wind fields. The angle (q) required for rotation of the axes into principal axes, and the resultant principal stretching deformation (np) are given by

q = 1
2
tan-1[(v1+u2)/(u1-v2)]     (3.5)
np = [(u1-v2)2+(v1+u2)2]1/2.
He also presents a method for streamline analysis of deformation and a method for evaluating deformation under the geostrophic approximation. The existence of a deformation field is a sufficient condition for frontogenesis, and dynamical studies have been carried out for the atmospheric fronts whose development is ultimately limited by secondary circulation (Stone, 1966) and dissipative processes using various approximate dynamical models. To discuss the rate of change of f moving with the fluid, use is made of the so-called frontogenesis function D/Dt|Ñf|2 (Hoskins, 1982). In certain cases, a simple shear normal to the flow direction has been implicated as a generation mechanism (Williams, 1967).

(c) Oceanic gradients

Relatively intense gradients occur in the deep and coastal oceans associated with a variety of scales of motion, phenomena, and internal and external driving mechanisms. Permanent fronts separate water masses and give rise to intense boundary and free jets. Permanent convergence zones exist in the equatorial, subtropical and polar oceans. Water mass driven, wind-driven, and tidally driven fronts occur at and across shelf breaks, in the coastal ocean and in the vicinity of banks. Internal dynamical processes (e.g., instabilities, nonlinear and topographic interactions) populate the ocean with frontal ring vortices, eddies, and lenses over a range of meso- and submeso-scales. Such fronts and eddies must be expected to produce biological fronts and interactions and to influence biological dynamics. Biological and physical interactions for some types of fronts, have been and are being studied, such as: eastern boundary upwelling systems (Barber and Smith, 1981), cold and warm core Gulf Stream rings (The Ring Group, 1981; Olson and Backus, 1985; Franks et al., 1986); tidal fronts (Lynch et al., 1995); and eddy-eddy interactions (McGillicuddy et al., 1995). Others have only recently been discovered or understood and await interdisciplinary research.

A theoretical study (Kirwan et al., 1994) of aspects of nonlinear ocean dynamics is based in part on the local kinematics of horizontal divergence (x), deformation (n) and vorticity (w). Waves eddies and submesoscale lens are considered under various dynamical assumptions. Further and broader studies based in the kinematic approach can be useful to the theory of advective effects in biological dynamics. A number of relevant field studies have been carried out attempting to measure directly the kinematic quantities from the dispersion of float clusters. Inference of x, n and w is quantitatively difficult and usually sensitive (Kirwan et al., 1988, 1992). A recent study (Sanderson, 1995) was carried out on the Western Bank of the Scotian Shelf in which the drifters were part of a measurement and modeling system to track cod larvae (Bowen et al., 1995). The estimates, considered sensitive, were

w = (-1.9±0.06)×10-5 s-1, x = (-9.5±7)×10-7 s-1,

n = (9.0±7)×10-7 s-1, m = (1.7±0.6)×10-6 s-1.

Recall that (1.16)×10-5s-1 = (day)-1. Thus such physical time scales are in the range of some important biological time scales (Denman and Powell (1984); Dickey (1990)). Some general scale considerations for effective biological and physical interactions are discussed by Steele (1988), Nihoul and Djendi (1991) and Denman (1992).


4. Biological processes


In the following section, problems have been encountered translating TeX constructs into HTML in text paragraphs. For example, the construct
^
a
 

20
in text becomes [^(a)]20. At present, this is unadvoidable and should be kept in mind while reading this document.

In this section we summarize some of the dynamical properties of a simple two-component biological system in zero spatial dimensions (e.g., Freedman (1980); Murray (1993)). We assume a predator-prey system with perfect grazing. In the absence of grazing, the prey can achieve an equilibrium value (carrying capacity) but the predator is depleted, e.g., a simple phytoplankton-zooplankton model with unlimited light and nutrient to support photosynthesis.

The relevant equations are (2.2) and (2.6) with nondimensionalization when desirable as in (2.10) but with u0 = 0. To keep all of the constants in the equations positive, we adopt the notation [^a]ij = -aij if aij < 0, and similarly for aij. The model equations under the above assumptions are, with a12 = -a21 and a22 = 0,

df1
dt
= f1(a10- ^
a
 

11 
f1-a21f2)     (4.1a)
df2
dt
= f2(- ^
a
 

20 
+ a21f1)     (4.1b)
We shall consider: the linearized stability properties of equilibria, the nonlinear approach to the carrying capacity equilibrium, and nonlinear grazing dynamics. Freedman (op. cit., Chapter 3) reviews properties of this model which he refers to as a Lotka-Volterra system with semidecreasing returns.

Let Fi denote a steady state equilibrium field and introduce the notation

f1 = F1+d,      f2 = F2+e   .     (4.2)
with (d,e) the general differences away from equilibrium. The equilibrium fields satisfy (4.1) with [d/ dt] = 0. Upon addition of (4.1a,b) and substitution of the result back into (4.1a), there results
F1(a10- ^
a
 

11 
F1)( ^
a
 

20 
-a21F1) = 0     (4.3)
F2 = F1
^
a
 

20 
(a10- ^
a
 

11 
F1)
There are three equilibrium states, viz.

F1 = 0,      F2 = 0            (null)     (4.4a)

F1 = a10
^
a
 

11 
º C,      F2 = 0            (carrying capacity)     (4.4b)

F1 =
^
a
 

20 

a21
º G,      F2 = G æ
ç
ç
ç
ç
è
a10
^
a
 

20 
-
^
a
 

11 

a21
ö
÷
÷
÷
÷
ø
            (grazing)     (4.4c)

(a) The null state

For instability analyses we assume d,e to be infinitesimal. Then linearization of (4.1a,b) about the null state (4.4a) decouples the two fields and f1 is unstable and f2 is stable. If the f2 perturbation is suppressed, f1 will evolve in time to the carrying capacity equilibrium C. If e ¹ 0, f2 will begin to grow when the second term on the rhs of (4.1b) exceeds the first term. Nondimensionalizing both state variables by M = C and choosing t0 = (a10)-1, eqs. (4.1) become (vid. (2.10c))

df1
dt
= f1(1-f1-a21f2)     (4.5a)
df2
dt
= f2(- ^
a
 

20 
+a21f1)     (4.5b)
where

^
a
 

20 
º
^
a
 

20 

a10
,      a21 º a21
a10
C = a21
^
a
 

11 
For a period of time after f1 begins exponential growth and achieves finite amplitude the departures from equilibrium will be governed by (4.5a) linearized around the null state (4.4a) and the full (4.5b), i.e.,
dd
dt
-d = 0     (4.6a)
de
dt
+e[ ^
a
 

20 
-a21d] = 0     (4.6b)
with solutions
d = d0et     (4.6c)
e = e0exp[- ^
a
 

20 
t+a21d0et].     (4.6d)
Thus at time tg determined by

tg = ln[
^
a
 

20 

a21
tg
d0
]   ,     (4.6e)
there will be an explosive growth of the predator f2.

(b) Carrying capacity

The carrying capacity equilibrium is governed by equations (4.5) with f2 º 0 and d = f1-1. The full d equation is

dd
dt
+d(1+d) = 0.     (4.7a)
The linearized equation now indicates stability
d = d0e-t     (4.7b)
and the nonlinear approach to equilibrium is given by the logistic growth solution
d = d0
(d0+1)et-d0
     (4.7c)
f1 = f10
f10+(1-f10)e-t
     (4.7d)
Figure 3a depicts f1(t); since the mortality is quadratic an excess of f1 over C is depleted more rapidly than a deficit is enhanced as the carrying capacity equilibrium is approached for all initial conditions.




Figure 3a

Figure 3. (a) Approach to carrying capacity




To study grazing we nondimensionalize the time by t0 = (a21M)-1, with M as an arbitrary biomass density. Equations (4.1) become

df1
dt
= f1(a10- ^
a
 

11 
f1-f2)     (4.8a)
df2
dt
= f2(- ^
a
 

20 
+f1),     (4.8b)
where

a10 = a10
a21M
= C
M
^
a
 

11 
,       ^
a
 

20 
=
^
a
 

20 

a21M
= G
M
,       ^
a
 

11 
=
^
a
 

11 

a21
.
To present the stability results for the grazing equilibrium (4.4c), we set M = G or [^(a)]20 = 1. Thus

F1 = 1;   F2 = a10- ^
a
 

11 
,   a10 > ^
a
 

11 
,     (4.9a)
and d,e (equations (4.2)) satisfy the linearized equations
d
dt
d = - ^
a
 

11 
d-e     (4.9b)
d
dt
e = (a10- ^
a
 

11 
)d,     (4.9c)
For solutions ~ exp[st],

2s
= - ^
a
 

11 
±[ ^
a
 
2
11 
-4(a10- ^
a
 

11 
)]1/2,    or
(4.9d)
= - ^
a
 

11 
±2(-a10)1/2    if          a10 >> ^
a
 

11 
(4.9e)
For [^(a)]11 º 0 this reduces to the neutrally stable oscillation of the classical Lotka-Volterra model with dimensional frequency [a10[^(a)]20]1/2. For all finite [^(a)]11 the grazing equilibrium is stable and for small [^(a)]11 the decay rate is given directly by [^(a)]11. The coefficients of the rhs of (4.9b,c) are the elements of the community matrix with trace and determinant

Tr = - ^
a
 

11 
< 0,      D = a10- ^
a
 

11 
> 0     (4.9f)
and (Tr)2 = 4D, when
^
a
 
*
1 
= 2(1+a10)1/2-2
For [^(a)]11 < a11* the grazing equilibrium is a stable node or stable spiral in phase space and [^(a)]11 = 0 is a neutral center. More generally, the grazing equilibrium has been shown to be globally stable in the Liapunov sense (Freedman, loc. cit.).

Next we present the solution for the nonlinear conversion of prey to predator for the limiting case that the grazing rate is much faster than growth and mortality rates. From equation (4.8) this limit is seen to require

M >> G,      M >> ^
a
 

11 
C,      1 >> ^
a
 

11 
     (4.10a)
Setting a10 = [^(a)]11 = [^(a)]20 = 0 in (4.8) yields solutions
f1 = f10(f20+f10)
f10+f20et
     (4.10b)
f2 = f20(f20+f10)et
f10+f20et
     (4.10c)
f1+f2 = f10+f20 = 1      (4.10d)
Asymptotically the total biomass (2.1) is in the predator f2. For specificity in the next section, we may regard f1-f2 as a phytoplankton-zooplankton system or as a nutrient-phytoplankton system. Nonlinear conversion is illustrated in figure 3(b).




Figure 3b

Figure 3.(b) nonlinear conversion [phi10 = .999, .8, .4];




The effect of mortality of the predator can be treated analytically via singular perturbation theory when the mortality rate is small but finite. Thus we keep a10 = [^(a)]11 = 0 but retain [^(a)]20 << 1 in (4.8) as the perturbation parameter. The solution (4.10b) for the prey is unaltered but the solution for the predator is now

f2 = (f20+f10) é
ê
ë
f20et
f10+f20et
+e-[^(a)]20t-1 ù
ú
û
.     (4.10e)
The solution is uniformly valid in time under the condition [^(a)]20 < f20, and has been checked by numerical integration for selected parameters. An example is plotted on figure 3(c).




Figure 3c

Figure 3. (c) effect of mortality on the predator after the prey is depleted;




Consider for example f1 as a chemical nutrient and f2 as phytoplankton. After consuming almost all the nutrient in the dimensional time scale (a21M)-1 the phytoplankton slowly die off on the dimensional time scale (a21G)-1, as shown on figure 3(c). The plankton f2 asymptotically approach zero identically but there is a small residual amount of nutrient f1¥ which can be computed from the exact integral of the system. Equation (4.10d) is now replaced by

f1+f2- ^
a
 

20 
lnf1 = f10+f20- ^
a
 

20 
lnf10 = 1-a20lnf10 .= 1     (4.10f)
Since f2¥ = 0 and f1¥ is small, we have
f1¥+f2¥- ^
a
 

20 
lnf1¥\doteq- ^
a
 

20 
lnf1¥\doteq 1     (4.10g)
whence

f1¥ = exp[-( ^
a
 

20 
)-1]   .     (4.10h)

(d) N-P-Z Model

The nonlinear conversion results are extended to a three component system. The phytoplankton f1 = P are grazed by the zooplankton f2 = Z while they consume the nutrient f3 = N. We assume perfect grazing (aij = -aji) and neglect mortalities (aio = aii = 0). The non dimensional model equations, with all biomasses scaled by M and t0 = (a13M)-1, are

Pt = PN-bPZ     (4.11a)
Zt = bPZ     (4.11b)
Nt = -PN     (4.11c)
This is a reduced form of the zero-dimensional NPZ model used by Klein and Steele (1985) as the basis for a numerical study of advective and diffusive effects in one and two spatial dimensions. They included nutrient recycling and uniform flow in idealized Georges Bank geometries.

The parameter b = a21 (a13)-1, the ratio of the grazing interaction coefficients, measures the relative rate of zooplankton depletion of phytoplankton to their growth rate via nutrient consumption. Addition of the three equations and an integration after eliminating P from (4.11b,c), yields two integrals of the system

N+P+Z = B     (4.12a)
ZNb = C     (4.12b)
and a quadrature solution
ó
õ
dN
[-N2 + BN- CN(1-b)]
= -t.     (4.12c)
Here B, C are constants of integration.

To facilitate the study of advective effects we seek relatively simple solutions to the zero-dimensional biological dynamics. Thus we will take the exact solution for b = 1 and approximate solutions for general b. For b = 1, (4.12c) integrates to

N = E( B- D)+( B+ D)e- Dt
2( E+e- Dt)
     (4.12d)
D = ( B2 - 4 C)1/2
which together with (4.12a, b) completes the solution. E is the third integration constant. In terms of the three initial values of biomass at t = 0,
B = N0 +P0 +Z0     (4.12e)
C = N0 Z0
E = B+ D-2N0
2N0 -( B- D)
.
A case of special interest is that in which the biomass is initially almost entirely in the nutrient with equal small background values of planktons. Then

N0 = O(1),       P0 = Z0 = e << 1  ,     (4.12f)
and

B = N0 +2e,       C = eN0,       D = N0,       E = e
N0
     (4.12g)
to accuracy O(e). Asymptotically in time

Z¥ = N0        N¥ = e      P¥ = e  .     (4.12h)
With equal grazing rates almost the entire nutrient supply is successfully converted into zooplankton. This solution is shown on figure 3(d) for N0 = 1.




Figure 3d

Figure 3. (d)  N,P,Z conversion.




If b = 0, the system reduces to the two components (N,P) and the solution becomes

N = N0(N0+e)e- Bt
N0e- Bt+e
,      P = e(N0+e)
N0e- Bt+e
,     (4.12i)

B = N0+e,      P¥ = N0,      N¥ = 0,
as for equations (4.10) and figure 3(b).

From (4.11a) we note that the initial rate of plankton growth is given by P0 (N0 - bZ0). Thus if b << 1, N0 will first be converted to P via the 2-component process described by equation (4.12i), followed by a similar two-component conversion of N0 to Z but on the longer time scale b-1. For b > N0 Z0-1, P decreases immediately at t = 0 and cannot efficiently act as a conduit for the growth of Z. For b >> N0 Z0-1 Z can only convert P0, and asymptotically N¥ = N0, P¥ = 0, Z¥ = Z0 + P0. Thus rapid grazing is disadvantageous for the zooplankton.

An interesting case lies in the determination of asymptotic values in the range N0 Z0-1 > b > 1; for conditions (4.12f) this becomes b < O(e-1). There is an initial growth of plankton to a maximum value Pm with associated values Zm, Nm, at t = tm. From (4.11a) Nm = bZm when Pt = 0. From (4.12a,b,f)

N+P+Z = N0 +2e,       N = N0 e[1/(b)] Z-[1/(b)]     (4.13a)
Zm = e[1/( 1+b)] b-[(b)/(1+b)] N0[(b)/( 1+b)]     (4.13b)
and Nm, Pm are given by (4.13a). After the maximum the asymptotic values are approached rapidly in the now faster time scale b-1. Inspection of (4.12a,b,c) shows P¥ = 0, and the exact asymptotics can be then computed from (4.13a).

(N0+2e-N¥)N¥b = eN0b,       Z¥ = N0+2e-N¥,     (4.13c)
or
N¥ = e1/bN01-1/b,      e << 1.
Given the exact values of the state variables at the maximum and at ¥, the solution can be completed approximately by using a two-component N-P conversion (4.12i) up to the maximum and a two-component P-Z conversion beyond the maximum. Table 1 presents some numerical values for the fields at tm and for the asymptotic partitioning of N0 between Z¥ and N¥.




Table 1. Field values at the maximum and at infinity

ebZmNmPmZ¥N¥
0.00105.00000.08270.41350.50580.75080.2512
0.001010.00000.06580.65790.27830.50080.5012
0.001015.00000.05130.76910.18160.37100.6310
0.0010100.00000.00980.97750.01480.06870.9333

0.0100

5.00000.12140.60700.29160.62190.3981
0.010010.00000.08110.81110.12780.38900.6310
0.010015.00000.05920.88820.07260.28440.7356




5. Interactive processes

In this section we consider the occurrence of the biological processes of the previous section in the presence of a dynamically interesting flow field with deformation. We will present results for the competition between physical and biological effects for: the stability analyses, the nonlinear approach to carrying capacity or advective equilibria, and nonlinear grazing.

The parameter a = u0t0L-1 (equation (2.10c)) measures the relative importance of advective to biological effects if t0 is chosen as the characteristic time of the dominant biological process. As for equations (3.4) we choose v1 = u2 = 0 in (3.2) and set u0L-1 = u1, the deformation gradient of the flow field y = -u1xy. Equations (2.10), (2.11), (2.12) in nondimensional form appear as

fi
t
+a é
ê
ë
x fi
x
-y fi
y
ù
ú
û
= Bi(fj)     (5.1a)

dt
ds
= 1       dx
ds
= ax       dy
ds
= -ay       df1
ds
= Bi     (5.1b)
a = u1t0,      ai0 = t0 ai0,      aij = Mt0 aij,    and
y = -xy
Note that because of the special form of the velocity field, L does not appear in a. The biological operator has been scaled by Mt0-1. Again we will treat the initial value problem in the time domain so the solutions (2.14b) are

s = t      r = xe-as      q = yeas.     (5.1c)
Since gradients in the direction of the axis of contraction (dilation) amplify (decay) it suffices to take the initial condition as a function of y only, i.e.,

fi(x,y,t = 0) = fi(q(y,0) = y,0)   ,     (5.1d)
which for a local gradient analysis will be as in (3.1a) with c1 = 0, i.e.,
fi(y,0) = g0i+g2iyeat     (5.1e)
g0i = c0i
M
,      g2i = c2iL
M
in nondimensional form.

(a) Linearized instabilities of equilibria

Since a > 0 always, q = yeat represents exponential growth, and the presence of an initial spatial gradient of fi can significantly alter stability properties in the presence of the physical flow. Stable solutions can be rendered unstable and instabilities will occur with enhanced growth rates.

Linearity allows separability, thus

fi = di(q)fi0(t) = (g0i+g2iyeat)fi0(t),     (5.2)
The fi0(t) are the zero-dimensional biological dynamical solutions for the null (n), carrying capacity (c), and grazing (g) states (equations (4.4)) whose stability properties are given by equations (4.6)-(4.7), (4.9). In those solutions, e.g. (4.6) the initial conditions d0, e0 are now replaced by d1(q), d2(q) from (5.2).

For the null state (4.6) the instability of the prey is enhanced and the stability of the predator can be switched to instability if the advective supply rate exceeds the decay rate. Sufficient advection can also destabilize both the carrying capacity (4.7) and grazing states (4.9); the Lotka-Volterra model in the presence of deformation advection is always switched from neutral stability to instability. The quantitative conditions among the biological rate and interaction parameters and the deformation shear are given in Table 2.




Table 2. Destabilizing shears

StateaConditionEffect
nullu1a10a > [^(a)]20 or  u1 > a20(a10)-2predator
unstable
carrying u1a10a > 1             u1 > (a10)-1instability
capacity
grazingu1[^a]202a > [^(a)]11            u1 > [^a]11(a21[^a]20)-1instability




Okubo (1980, Chapter 10) reports a stabilization of the Lotka-Volterra model when the predator and prey tend to migrate towards a central favorable region with constant velocities in the presence also of diffusion.

(b) Carrying capacity vs. advective equilibria

Here we consider the behavior of the one component biological system, e.g., phytoplankton (4.7), in the presence of the deformation advection of equation (5.1). The solutions are now given by (4.7c,d) with d0, f10 replaced by functions of q as in (5.1c).

First we generalize the stability analysis of equation (5.2) for case g0 = 0, g2 = 1. Now M = C and L = Cc2-1. The nonlinear solution for the departure from equilibrium is

d = d
(d+1)et-d
,      
y > y0 = -e-at
(5.3a)
d = -1                  
y < y0 = -e-at
d = yeat     (5.3b)
The initial (t = 0) population is null for y < -1 and varies linearly with y for y > -1, passing through the carrying capacity equilibrium value at y = 0.

The limit of pure advection is given by d = d directly by (5.3c) or equivalently by setting et = 1 in (5.3a) but retaining eat. The limit of pure biology is given by a = 0. Both of these limits are plotted on figure 4(a). Advection is plotted for a = 1, i.e., with equal advective and biological time scales. To interpret these and subsequent plots note that at t = 0, d = y, so that a curve emanating from a point d* on the ordinate axis represents d(y = d*,t). The biological behavior is of course identical to that shown on figure 3(a). The advective solutions for y > 0 amplify indefinitely but for y < 0 reach the limiting value d = -1 at time t-1 = ln(-y-1) and remain at that value thereafter. This behavior is consistent with the deformation flow (vid. figure 2(b)) ultimately advecting in fluid with no phytoplankton for y < 0 but advecting in fluid with an ever increasingly dense population for y > 0.




Figure 4a

Figure 4a

Figure 4. Carrying capacity vs. advective equilibria: (a (above) ) Pure advection (dashed) and pure biology (solid); (b (below) ) dependance on relative strength of advection [Click on components of 4b for increased resolution].


Alpha = 0.1 Alpha = 1.0 Alpha = 10.0

Figure 4b




The general interactive solution is illustrated on figure 4(b) for values of a = (0.1,1,10) which range from relatively weak to strong advective effects and span all qualitative behaviors. The solutions equation (5.3a) have maxima at t = tm for both positive and negative y but only for a range of values. The times at which maxima and limiting values are reached are

tm = 1
a
ln( a-1
y
)

y > 0,   a > 1,    y < a-1
y < 0,   a < 1,    y < a-1
     (5.3c)
and
t-1 = 1
a
ln æ
ç
è
- 1
y
ö
÷
ø
   .

Asymptotically in time an excess initial population always approaches the carrying capacity equilibrium but a deficit population approaches the advective null equilibrium.

Examination of figure 4(b) shows some interesting effects. Frontogenesis occurs, ultimately producing a very sharp front of unit strength. For large a and positive y, advection dominates until the maximum is reached at tm and a much stronger but broader front is present. For large a and negative y advection dominates with but little biological modification before t-1. For small a advection has little effect on the approach to the biological equilibrium for positive y. For negative y and small a, however, the solutions are dominated by almost pure biological behavior for a long time, i.e., until tm >> 1. Biological equilibrium is very nearly achieved whence advection suddenly sweeps away the population. This is because (equations (5.3)) for a << 1, tm » t-1. The curve emanating at y = -.20 for a = 0.10 which is indistinguishable from zero at t = 8 has the associated values tm = 15.04, t1 = 16.09.

The next problem deals with the identical dynamics as the previous one but for the initial condition introduces a finite frontal width L across which the biomass density changes from zero to a maximum value of F. We retain the biomass scaling M = C as introduced prior to equation (5.3a). The initial condition is now taken as

f1(y,0) = g(1+tanhy),      g = F
2C
     (5.4a)
f1(y,0) = g(1+y),       y << 1

f1(¥, 0) = 2g,   f1(0,0) = g,   f1(-¥,0) = 0.
The cross frontal coordinate has been scaled by L, recall that tanh(2) = 0.96. Note that in the vicinity of the y-origin this solution will correspond to that of the preceding section if g = 1.

The approach to equilibrium is now given by

f1 = d1
d1+(1-d1)e-t
= 1
1+(1/d1-1)e-t
     (5.4b)

d1 = g(1+tanhyeat)   .     (5.4c)
Since d1 remains finite for all time for y ³ 0, the asymptotic behavior is a simple exponential decay in time to the biological equilibrium f1 = 1. For negative y, d1 approaches zero as

f1 = 2gexp(2yeat+t)       y < 0,     t®-¥.     (5.4c)
The interpretation, illustrated below, is that for y < 0 the asymptotic limit is always a decay to the advective null equilibrium, but that before a time ta determined by
tae-ata = -2y     (5.4d)
there can be a growth towards the biological equilibrium unity, which can be approached very closely for a < 1. Note that for g < 0.5 the initial condition is everywhere a deficit of population density from the carrying capacity. As advection tightens the front across y = 0, the excess or deficit of density advected in from the populated side of the front is adjusted by biological processes to the carrying capacity.

The range of qualitative behavior of solution (5.4) as a function of the relative rate and frontal strength parameters (a,g) is illustrated on figure 5. Figure 5(a) presents the pure biological and pure advective limits. Figures 5(b,c) are for an initial frontal strength (2g) somewhat larger than the carrying capacity and Figures 5(d,e) show the behavior for a weaker (stronger) front. Figure 5(c) illustrates a small a behavior analogous to that on figure 4(b).




Figure 5a

Figure 5a


Alpha = 0.1 Alpha = 1.0 Alpha = 10.0

Figure 5b (click on components for increased resolution)


Alpha = 0.1 Alpha = 1.0 Alpha = 10.0

Figure 5c (click on components for increased resolution)


Alpha = 0.1 Alpha = 1.0 Alpha = 10.0

Figure 5d (click on components for increased resolution)

Alpha = 0.1 Alpha = 1.0 Alpha = 10.0

Figure 5e (click on components for increased resolution)

Figure 5. As for Fig. 4 but for a finite front: (a) Pure advection (dashed) and pure biology (solid); (b) and (c) initial frontal strength somewhat larger than carrying capacity and show, respectively, the short and long time behavior; (d) initial frontal strength smaller than carrying capacity; (e) initial frontal strength much larger than carrying capacity.




(c)Grazing with advection

The effects of advection on grazing are particularly interesting since advective supply of nutrients can dominate the qualitative and quantitative behavior of the biological dynamics. We retain the deformation flow field of equation (5.1). For definiteness of interpretation, we can consider solutions in the half-plane y > 0 as representing an infinitely deep ocean with sea surface at y = 0. Thus advection can be regarded as bringing material from the deep sea into the surface region. The models remain of course highly ideaized and we do not intend to imply any direct realistic oceanic relevance.

The advective process introduces a new advective equilibrium state. Introducing the biology of equations (4.10) into equations (5.1a,b) and setting [()/(t)] = [()/(x)] = 0 yields

-ay P
y
= PN     (5.5a)
-ay N
y
= -PN,      (5.5b)
with solutions
N = B
1+ Ae-[( B)/(a)]lny
     (5.5c)
P = ABe-[( B)/(a)]lny
1+ Ae-[( B)/(a)]lny
     (5.5d)
N+P = B     (5.5e)
where A, B are pure constants of integration. B is the biomass at y = ¥, where P = 0. When y = 0, P = B and N = 0; when y = 1, N= B(1+ A)-1. The structure of the nutricline can be seen on figure 6 to depend upon A but to be primarily controlled by a. The advecting velocity, the source at infinity, and the geometry require a specific spatial distribution of biological fields. Advective equilibria are an important aspect of physical-biological interactions in the sea, but are not the focus of this study. We mention the solution (5.5) for completeness but return to our central theme which is to introduce deformation flow advection into the nonlinear grazing dynamics of equations (4.10), (4.12), (4.13).




Figure 6

Figure 6. Advective equilibrium profiles [alpha = .1 (solid), 1 (dash-dot), 10 (dashed)]. The A dependence is indicated by where the alpha triplets intersect at y = 1. For A = (.1,1,10), the intersections at y = 1 are at N = (.9,.5,.09), respectively.




We need consider only equations (4.11)-(4.13) since they contain the dynamics of (4.10) for the case of b = 0. For all solutions we will take the initial plankton fields as a small uniform background constant. The initial nutrient field will be a linear increase of N with depth y, and interest will be restricted to order unity depths. Thus

N(0) = 1+y,       Z(0) = P(0) = e     (5.6a)
and the advective solutions are given by (4.12), (4.13) with
N0(r,q) = 1+q = 1+yeat,     (5.6b)
N0 = yeat,       t®¥
from (5.1c,d). The asymptotic values for the two and three component systems are now given by (4.12i,h) and (4.13c) with N0 given by (5.6b).

For the two-component NP system figure 7 illustrates the dependence of the grazing on the initial condition (e) and the advective parameter (a). Figure 7(a) shows an initial advective growth of P followed by a complete depletion via grazing. It is important to note that after some times the solution is independent of the value of the initial phytoplankton biomass, which simply controls the amplitude and time of occurrence of the maximum of N. Figures 7(b,c) for fixed e = .001 illustrates the dependence on the advection rate. Rapid advection results in an early, large amplitude maximum in N and an early rapid growth of P asymptotically.




Figure 7a

Figure 7a


Figure 7b

Figure 7b


Figure 7c

Figure 7c

Figure 7. NP conversion: (a) Dependence on initial conditions; (b) weak and moderate advection [N = solid, P = dashed]; (c) moderate and strong advection [alpha = 1.0 (solid), 2.5 (dash-dot), 10 (dashed)], as a increases, the maximum in N increases and the conversion to P occurs earlier; (d) grazing and advection rates [y = 0.5 (dashed), 1.0 (solid)], as time increases local grazing stops.




Recall that if there were no planktons (Z = P º 0) the purely advective solution for the inert passive nutrient would be given by (5.6b) (vid equation (3.4b)). In the advective NP system (b = 0) for large t

P¥ = N0 = yeat,       N¥ = 0     (5.6c)
indicate that the phytoplankton effectively utilize all of the advected nutrient and continue to grow and accumulate an indefinitely large biomass in the upper ocean with no observable nutrient present. The local time derivative (equations (4.12i, 5.6b))
æ
ç
è
P
t
ö
÷
ø


x,y 
= æ
ç
è
P
t
ö
÷
ø


N0 
+ æ
ç
è
P
N0
ö
÷
ø


t 
æ
ç
è
N0
t
ö
÷
ø
= NP+ay P
y
= G+A
(5.6d)
= eN0(N0+e)2eBt
(N0+eeBt)2
+ eeBt
(N0+eeBt)2
[e(eBt-1)+tN0(N0+e)]ayeat
has a contribution from local grazing, G, and advective flux, A. These contributions are plotted vs. time on figure 7(d). At a given depth (y) local grazing has a maximum and then decays to zero; advective flux peaks, decays and then grows asymptotically. At greater depths, the extrema occur earlier and the maxima are larger. Thus the asymptotic growth of Z at a given depth is due to an advective flux from a grazing conversion which occurred earlier and deeper. The character of the solutions at very large times and depths is only of mathematical interest. Biological interpretations should be restricted to finite ranges of (y,t).




Figure 7d

Figure 7d




The behavior of the three-component NPZ system for equal grazing rates (b = 1) is shown on figures 8(a,b) for the same values of e and a and the same scales as figures 7(b,c) for comparison. As indicated by equation (4.12h), the advected nutrient is now asymptotically completely converted to Z with ultimately no observable N or P. The rate of the processes again increases with increasing a, as do the amplitudes of the successive maximum values of N and P. For b >> 1, 1 ³ eb, the fraction of nutrient effectively utilized by P and Z and its partioning between them, is given by (4.13c). N and Z will now both be observed to grow exponentially after the initial adjustment time. In this model, rapid grazing by zooplankton leads to a smaller Z biomass than slow grazing.




Figure 8a

Figure 8a


Figure 8b

Figure 8b

Figure 8. NPZ conversions: (a) Weak and moderate advection [N = solid, P = dashed, Z = dash-dot]; (b) moderate and strong advection, maxima in N and P increase with increasing a and the rapid growth of Z occurs earlier.




6. Summary and conclusions

This study presents a general theoretical structure for the analysis of advective effects on biological dynamics in the sea. It is based upon the mathematical theory of characteristics and, in the domain of the parameter s, involves the zero-dimensional local biological dynamics, as well as the Lagrangian trajectories of the fluid particles. The s-domain initial value problems solved here were for the case that s was identified with the time t, and involved the specification of space-wise initial fields at t = 0. More general problems are possible and should be considered.

A kinematical analysis of the flow field was introduced and the effects of divergence, vorticity, stretching and shearing deformations on local gradients of biological state variables were discussed. The stretching deformation flow, which contains frontogenesis and frontolysis effects that contract and dilate gradients, was selected for the present study. The biological dynamical system included net birth or mortality, self-interactions and coupled field interactions. Biological processes subjected to advection included: the approach to local equilibria; the stability or instability of equilibrium states for one, two, or three-component systems; and nonlinear grazing conversions for NP and NPZ models. Further research should be directed towards additional and more realistic flow structures (vorticity, divergence, three dimensions, time-dependence, and nonlocal circulation structures, etc.) as well as additional and more realistic biology (Michaelis-Menten grazing, light limitation, behavior, etc.).

Advection by a deformation shear field in the presence of local gradients of the biological state variables was found capable of destabilizing null, carry-capacity and grazing equilibria and the velocity shear required was obtained in terms of the biological growth (death) rates and interaction coefficients. The competition between local biological and advective equilibria in the presence of frontogenesis was illustrated both for a simple deformation field flow and for a front of finite width. For an NPZ model the concept of advective-grazing equilibria was introduced and exemplified. For the cases of initial small background distribution of planktons in the presence of a deep ocean reservoir of nutrient the growth and conversion of planktons were calculated including the asymptotic in time partitioning of the advective nutrient input among the NPZ variables. At large times, P does not survive and Z is most successful by slow rather than rapid grazing.

These first results indicate the utility of the approach and provide some insight for further research. Deformation fields occur over a range of scales in the ocean and in other fluids in which dynamic tracers are of interest for a variety of reaction and nonconservative dynamics. Biological rates and physical advection times of interest in the ocean range from hours to days to weeks to months and seasons. A comprehensive review of time scales in the context of this theory could serve both to constrain the theory and guide developments. In addition to the direct extensions discussed in the preceding paragraphs it would be of interest to carry out first studies including new processes such as chaotic trajectories, mixing, swimming and small-scale encounter dynamics. The formulation of more general initial value problems in the s-domain should be guided not only by direct biological considerations but also by the mathematical structure of the theory of characteristics.



Acknowledgements

This research was stimulated by several interesting, important and enjoyable scientific discussions with Prof. Brian J. Rothschild. I thank Profs. A. Denny Kirwan, Jr. and Eileen E. Hofmann and Drs. Patrick J. Haley, Jr. and Dennis J. McGillicuddy for helpful scientific and mathematical comments. I am grateful to Mr. Wayne G. Leslie for his essential contribution to this study by the performance of sensitivity analyses and the preparation of figures. It is a pleasure to thank Dr. Nadia Pinardi for the hospitality of Istituto per lo studio delle Metodologie Geofisiche Ambientali (CNR-IMGA) during a sabbatical visit. The preparation of the manuscript by Ms. Marsha A. Glass and Mrs. Renate D'Arcangelo is very much appreciated. This research was supported in part by the Office of Naval Research under grant N00014-95-1-0371 to Harvard University.

References



Barber, R.T. & Smith, R.L. 1981 Coastal and estuarine fronts. In Analysis of Marine Ecosystems (ed. A.R. Longhurt). New York: Academic Press.

Batchelor, G.K. 1967 An introduction to fluid dynamics. Cambridge: Cambridge University Press.

Bergeron, T. 1928 Über die dreidimensionale verknüpfende Wetteranalyse I. Geofis. Pub. 5(6), 1-111.

Bowen, A.J., Griffin, D.A., Hazen, D.G., Matheson, S.A. & Thompson, K.R. 1995 Ship-board nowcasting of shelf circulation. Cont. Shelf Res. 15, 115-128.

Courant, R. & Hilbert, D. 1989 Methods of mathematical physics. In Partial Differential Equations, vol. II, 830pp. New York: Wiley Interscience.

Denam, K.L. & Powell, T.M. 1984 Effects of physical processes on planktonic ecosystems in the coastal ocean. Oceanogr. Mar. Biol. Annu. Rev. 22, 125-168.

Denman, K. 1992 Scale determining biological/physical interactions in oceanic food webs. In Aquatic Ecology; Scale, Pattern and Process (ed. P.S. Giller, A.G. Hildrew & D.G. Raffaelli). Oxford: Blackwell.

Dewar, W.K. & Flierl, G.R. 1985 Particle trajectories and simple models of transport in coherent vortices. Dyn. Atmos. Oceans 9, 215-252.

Dickey, T.D. 1990 Physical-optical-biological scales relevant to recruitment in large marine ecosystems. In Large Marine Ecosystems: Patterns, Processes, and Yields (ed. K. Sherman, L.M. Alexander & B.D. Gold), Publ. no. 90-30s, pp. 82-98. Washington, DC: Am. Assoc. Advanc. Sci.

Flierl, G.R. 1981 Particle motions in large amplitude wave fields. Geophys. Astrophys. Fluid Dyn. 18, 39-74.

Franks, P.J.S., Wroblewski, J.S. & and Flierl, G.R. 1986 Prediction of phytoplankton growth in response to the frictional decay of a warm-core ring. J. Geophys. Res. 91(C6), 7603-7610.

Freedman, H.I. 1980 Deterministic models in population ecology, 254pp. New York and Basel: Marcel Dekker, Inc.

Hinze, J.O. 1975 Turbulence. New York: McGraw-Hill, 790pp.

Horne, J.K. & D.C. Schneider 1994 Analysis of scale-dependent processes with dimensionless ratios. OIKOS 70, 201-211.

Hoskins, B.J. 1982 The mathematical theory of frontogenesis. Annu. Rev. Fluid Mech. 14, 131-151.

Kirwan, A.D., Jr. 1988 Notes on the cluster method for interpretating relative motions. J. Geophys. Res. 93(C8), 9337-9339.

Kirwan, A.D., Jr., Lewis, J.K., Indest, A.W., Reinersman, P. & Quintero, I. 1988. Observed and simulated properties of loop current rings. J. Geophys. Res. 93(C2), 1189-1198.

Kirwan, A.D., Jr., Indest, A.W., Liu, J. & Clark, N. 1990 Ring evolution in general circulation models from path analysis, J. Geophys. Res. 95(C10), 18,057-18,073.

Kirwan, A.D., Jr., Lipphardt, Jr., B.L. & and Liu, J. 1992 Negative potential vorticity lenses. Int. J. Eng. Sci. 30(10), 1361-1378.

Kirwan, A.D., Jr., Lipphardt, B.L., Jr. & Gregory, K.L. 1994 Nonlinear ocean dynamics. In The Oceans: Physical-Chemical Dynamics and Human Impact (ed. S.K. Majumdar, E.W. Miller, G.S. Forbes, R.F. Schmalz & A.A. Panah). Easton, PA: The Pennsylvania Academy of Science, pp. 40-55.

Klein, P. & Steele, J.H. 1985 Some physical factors affecting ecosystems. J. Mar. Res. 43, 337-350.

Lamb, H. 1932 Hydrodynamics. Cambridge: Cambridge University Press.

Lynch, D.R., Ip, J.T.C., Naimie, C.E. & Werner, F.E. 1995 Convergence studies of tidally-rectified circulation on Georges Bank. In Quantitative Skill Assessment for Coastal Ocean Models (ed. D.R. Lynch & A.M. Davies), Coastal and Estuarine Studies, Am. Geophys. Union 47, pp. 153-174.

McGillicuddy, D.J., Robinson, A.R. & McCarthy, J.J. 1995 Coupled physical and biological modelling of the spring bloom in the North Atlantic (II): Three-dimensional bloom and post-bloom processes. Deep Sea Res. I 42(8), 1359-1398.

Murray, J.D. 1993 Mathematical biology. Berlin: Springer-Verlag, 767pp.

Nihoul, J.C.J. & Djenidi, S. 1991 Hierarchy and scales in marine ecohydrodynamics. Earth-Sci. Rev. 31, 255-277.

O'Brien, J.J. & Wroblewski, J.S. 1973 On advection in phytoplankton models. J. Theor. Biol. 38, 197-202.

Okubo, A. 1980 Diffusion and ecological problems: Mathematical models, 254pp. New York: Springer-Verlag.

Olson, D.B. & Backus, R.H. 1995 The concentrating of organisms at fronts: a cold-water fish and a warm-core Gulf Stream ring. J. Mar. Res. 43, 113-137.

Ottino, J.M. 1989 The kinematics of mixing: Stretching, chaos, and transport. Cambridge: Cambridge Univ. Press.

Ottino, J.M. 1990 Mixing, chaotic advection, and turbulence. Annu. Rev. Fluid Mech. 22, 207-253.

Petterson, S. 1935 Contribution to the theory of frontogenesis. Geofis. Pub. 11(6).

Platt, T., Denman, K.L. & Jascy, A.D. 1977 Modeling the productivity of phytoplankton. In Marine Modeling, The Sea (ed. E.D. Goldberg, I.N. McCave, J.J. O'Brien, & J.H. Steele), vol. 6, pp. 807-856. New York: J. Wiley & Sons.

Platt, T. & Denman, K.L. 1975 A general equation for the mesoscale distribution of phytoplankton in the sea. Mem. Soc. of Sci. Liege 7, 31-42.

Pond, S. & Pickard, G.L. 1978 Introductory Dynamical Oceanography. Oxford, England: Pergamon Press, 241pp.

Ring Group, The 1981 Gulf Stream cold-core rings: their physics, chemistry, and biology. Science 212, 1091-1100.

Sanderson, B.G. 1995 Structure of an eddy measured with drifters. J. Geophys. Res. 100(C4), 6761-6776.

Saucier, W.J. 1953 Horizontal deformation in atmospheric motion. Transact. Am. Geophys. Union 34(5), 709-719.

Simpson, J.H. & James, I.D. 1986 Coastal and estuarine fronts. In Baroclinic Processes on Continental Shelves (ed. C.N.K. Mooers). Washington, DC: Am. Geophys. Union, pp. 63-93.

Steele, J.H. 1988 Scale selection for biodynamic theories. In Toward a Theory on Biological Physical Interactions in the World Ocean (ed. B.J. Rothschild). Dordrecht, The Netherlands: Kluwer Academic Publishers, pp. 513-526.

Stewart, H.J. 1945 Kinematics and dynamics of fluid flow. In Handbook of Meteorology (ed. F.A. Berry, Jr., E. Bollay, & N.R. Beers), pp. 412-419. New York: McGraw-Hill.

Stone, P.H. 1966 Frontogenesis by horizontal wind deformation fields. J. Atmos. Sci. 23(5), 455-465.

Williams, R.T. 1967 Atmospheric frontogenesis: A numerical experiment. J. Atmos. Sci. 24, 627-641.

Wroblewski, J.S. 1983 The role of modeling in biological oceanography. Ocean Sci. and Eng. 8(3), 245-285.








File translated from TEX by TTH, version 1.65.