Chapter 6
Time Dependence in the 1-D Models

In Chapter 5, we found that time-independent, zonally symmetric models can be cast into the form of a linear differential equation in the latitude variable with zero-flux boundary conditions at the poles. In the case of a latitude-independent thermal diffusion coefficient, the system can be solved with Legendre polynomials. In this chapter, we extend this procedure to the case of time dependence. Such an extension allows for idealized seasonal-cycle simulations for the zonally symmetric planet. Introducing time dependence necessarily requires some kind of effective heat capacity for the column of air–land or air–ocean medium. When we take the planet to have a homogeneous surface and constant diffusion coefficient, we also specify that the heat capacity is not spatially dependent, that is, it is a constant. An all-land planet is one for which the heat capacity is characteristic of a land surface. This means it takes a value calculated from a fraction (nominally half) of the mass of a column of air. This leads to a time constant c06-math-001 of about 1 month. This is our new phenomenological coefficient for Chapter 6.

The time-dependent models are amenable to analytical solution for uniform planets (constant c06-math-002 and c06-math-003). Model solutions are in the form again of Legendre polynomial modes, each having a characteristic decay time with larger scales (low Legendre index) having longer relaxation times, and smaller scales, shorter times. The seasonal cycle of latitude-dependent insolation has a very simple form when truncated at the few-mode level. Legendre polynomials of index 0, 1, and 2 describe the latitude dependence of the seasonal cycle of insolation, except, of course, for the discontinuous derivative near the poles, which is necessary to describe the onset of perpetual night or day and the appropriate latitudes and times of year. This insolation formula allows us to obtain very simple expressions for the dependence of insolation on obliquity, precession of the equinoxes, and eccentricity. The seasonal cycle simulations also give us an idea about the lag of seasonal response to the driving seasonal heating cycle.

Later in the chapter, we return to the stochastic nature of climate forcing (presumably due to weather instabilities) by showing how the random walk of heated parcels is approximately imitated by diffusion in an ensemble average. This analysis relates the mean square of the parcel's velocity and relaxation time to the diffusion coefficient of the ensemble.

A short section is added to show how one might go about solving one-dimensional problems via finite difference methods. The appendices to the chapter givederivations of the insolation functions.

6.1 Differential Equation for Time Dependence

Consider the problem of a time-dependent one-dimensional climate system for a planet whose surface is uniformly land or ocean. The energy balance model for constant coefficients is as follows:

6.1 equation

where c06-math-005 is the c06-math-006(latitude), c06-math-007 is time, c06-math-008 is the latitude and time-dependent temperature field, c06-math-009 is the insolation function possibly allowing for a seasonal cycle, c06-math-010 is the coalbedo, which is a function of latitude and possibly time dependent; c06-math-011 is the total solar irradiance divided by 4 = c06-math-012, c06-math-013, and c06-math-014 are phenomenological coefficients as defined in Chapter 5, and c06-math-015 is an effective heat capacity. As discussed in Chapter 2, this means a timescale of about 30 days (c06-math-016) for an all-land planet and a few years for a planet covered with a mixed-layer ocean. This amounts to using the heat capacity of half a column of air (at constant pressure) for the all-land case and that of 50–100 m of water for the mixed-layer ocean case. In the latter, we assume the ocean below is uncoupled with the mixed layer.

We expand the time-dependent temperature field into Legendre polynomials:

6.2 equation

Note that each coefficient c06-math-018 is a function of c06-math-019. After inserting the series, multiplying through by c06-math-020 and integrating as before we obtain an ordinary differential equation for the c06-math-021:

where

6.4 equation

It is important to note that none of the terms in the array indicated by (6.3) is coupled to any of the others. In other words, for the equation dependent on mode number c06-math-024 there is no other mode index, say c06-math-025, that is referred to in the equation indexed by c06-math-026.

6.2 Decay of Anomalies

Consider first the case where c06-math-027 and c06-math-028 are set at their mean annual values as in Chapter 5. We can examine the behavior of the climate as it is perturbed away from steady state. We have the solution to the initial-value problem:

6.5 equation

where c06-math-030 is the solution to the time-independent (steady-state) problem given above and

6.6 equation

Each mode c06-math-032 has its own characteristic time c06-math-033. Note that the characteristic times fall off rapidly as a function of c06-math-034 as shown in Figure 6.1. This agrees with our intuition that larger spatial scales have longer characteristic (adjustment) times. We will establish a similar result for the autocorrelation times later in noise-driven models.

Image described by caption and surrounding text.

Figure 6.1 This figure illustrates that large spatial scales (low Legendre index) have longer time constants than for small spatial scales. Shown is a bar chart with relaxation times for various Legendre modes for the homogeneous planet with c06-math-035 in terms of c06-math-036 month. In this case, c06-math-037 W mc06-math-038C, c06-math-039 W mc06-math-040C, so that c06-math-041 (dimensionless).

Image described by caption and surrounding text.

Figure 6.2 Decay of a symmetric (between the hemispheres) initial distribution (solid line) after times 0.25c06-math-042, 0.5c06-math-043, 0.75c06-math-044 and c06-math-045 (dashed lines). The model parameters are as earlier and the sum is truncated at c06-math-046. The steady-state response has a discontinuous derivative at the ring's latitude. Its Fourier–Legendre series is very slowly converging, and this indicates that much power resides in higher-mode indices. These modes decay quickly leaving behind a smoother latitudinal profile as the temperature relaxes toward the old steady-state solution (zero here).

6.2.1 Decay of an Arbitrary Anomaly

An arbitrary distribution of thermal anomaly will decay in time according to a superposition of the modes just discussed. For example, an initial distribution of anomaly (departure from equilibrium) whose shape is c06-math-047 will decay according to

For example, consider an initial shape rendered by a steady ring of heat source (Chapter 5) at a specific latitude as given by (5.11) and shown in Figure 5.11. The solid line in Figure 6.2 shows such a steady-state anomaly located at c06-math-049 (37c06-math-050N). If suddenly the heat source is switched off, the distribution will decay according to (6.7). The figure shows stages of the decay at 0.25c06-math-051, 0.5c06-math-052, 0.75c06-math-053, and 1.00c06-math-054. We see that the higher modes, those rendering the cusp-like peak, are lost quickly leaving behind only the smoother long-lived modes.

6.3 Seasonal Cycle on a Homogeneous Planet

We can begin the study of the seasonal cycle by placing the solar heating distribution as a forcing on the right-hand side of the energy-balance equation. The heating distribution can be written for a circular orbit as follows:

6.8 equation

where c06-math-056, and c06-math-057 (see Figure 6.3). The subscripts denote the Legendre index first, then seasonal time harmonic second. A derivation for this heating distribution function is sketched in the appendix to this chapter. But for now we can observe some important properties. The mean annual version can be quickly recovered by averaging from 0 to 1 year in c06-math-058,

6.9 equation

which is the formula used earlier in Chapter 5. The main part of the seasonal cycle is carried by the coefficient of the c06-math-060 term. Note that it is antisymmetric between the hemispheres (recall the identity c06-math-061). The seasonal forcing is largest near the poles (c06-math-062). Another interesting feature captured by this representation is that along the equator (c06-math-063) there are two maxima (see Figure 6.3). This is the effect of the Sun crossing the equator twice each year at the equinoxes. Table A.6.1 shows the coefficients for some higher terms in the expansion for the present elliptical orbit of the Earth. Figure 6.4 shows the seasonal insolation when only modes 0, 1, and 2 are included. This figure indicates that retention of only these three modes captures the most important features of the insolation. Figure 6.5 shows how in the tropics there are two peaks as one passes through the year. The figure shows that for c06-math-064 the semiannual harmonic is strong. These three modes are able even to capture the passage of the Sun over the equator twice per year. Away from the equator, the seasonal harmonic dominates. Figure 6.6 shows latitudinal time sections of the forcing. The dashed lines show the forcing for zero eccentricity.

Image described by caption and surrounding text.

Figure 6.3 Contour diagrams of the seasonal forcing c06-math-065 or insolation. The vertical axis is cosine of colatitude, c06-math-066; the horizontal axis is time, c06-math-067 in years and c06-math-068 corresponds to summer solstice for the Northern Hemisphere. The units of the contour plot is for the present elliptical orbit forcing through the Legendre mode 4 and time harmonic 2. The seasonal cycle of c06-math-069 for the present elliptical orbit with eccentricity 0.016 (a) and the seasonal cycle for the same heating, but for a circular orbit (b). The value of time equal to zero in the Northern Hemisphere winter solstice.

Image described by caption and surrounding text.

Figure 6.4 Seasonal cycle of c06-math-070 for the present orbital obliquity c06-math-071 when only the Legendre modes 0, 1, and 2 are retained along with mean annual, annual harmonic, and semiannual harmonic. The time span is over 2 years in units of years with the origin at northern winter solstice. The latitudinal coordinate c06-math-072 runs from the South pole (c06-math-073) to the North Pole (c06-math-074). The semiannual harmonic captures the passage of the Sun over the equator twice a year in this image.

Image described by caption and surrounding text.

Figure 6.5 Seasonal cycle of c06-math-075 in the tropics (c06-math-076) for the present orbital obliquity c06-math-077 when only the Legendre modes 0, 1, and 2 are retained along with mean annual, annual harmonic, and semiannual harmonic. This graphic shows the tropical variation with two maxima per year caused by the c06-math-078 term. Note the change in vertical scale from the previous figure. The time span is over 2 years in units of years with the origin at northern winter solstice. Note that the details of polar day and night are missed in this truncation, but the semiannual harmonic captures the passage of the Sun over the equator twice a year.

Image described by caption and surrounding text.

Figure 6.6 This figure illustrates the effect of eccentricity on the seasonal cycle of zonal average temperatures for an all-land planet. Shown is the modeled seasonal cycle (retaining Legendre modes 0, 1, and 2) of temperatures for the bare planet for present orbital parameters (solid curves) at selected latitudes: c06-math-0790.6 (36.9N/S), c06-math-0800.4 (23.6N/S), c06-math-0810.2 (11.5N/S), c06-math-0820.1 (5.7N/S), and 0 (Equator). The dashed curves are for the corresponding circular orbit (c06-math-083).

We can express the seasonal dependence of insolation as

6.10 equation

If we take the coalbedo, diffusion coefficient, and heat capacity to be constants independent of season and latitude, we can write equations for the mode responses:

6.11 equation

where c06-math-086 is the constant coalbedo. To simplify the algebra. we employ complex notation (superscript c06-math-087 indicates a complex variable) and the complex Fourier series:

6.12 equation

Inserting this into the governing mode equations:

6.13 equation

To recover the physical mode amplitudes,

6.14 equation

whereas the phase lag behind the forcing is

6.15 equation

The complex c06-math-092 are related to the real Fourier coefficients as follows:

6.16 equation

After a time (c06-math-094), the transients die out and we are left with the repeating steady-state solutions. First, consider the time-independent parts:

6.17 equation
6.18 equation

Next, consider the seasonal harmonic terms:

6.19 equation
6.20 equation
6.21 equation

and finally, the semiannual terms:

6.22 equation
6.23 equation
6.24 equation

In each case, we can compute the amplitude

6.25 equation

with phase lag

6.26 equation

It is interesting to evaluate these quantities for the all-land planet. Using typical values of the parameters, c06-math-100 W mc06-math-101, c06-math-102W (mc06-math-103 Kc06-math-104), c06-math-105, c06-math-106, we find c06-math-107 C, c06-math-108 K, c06-math-109 days, c06-math-110 K, c06-math-111 K, c06-math-112 days. While the phase lag is approximately correct for an all-land planet, the amplitude of both harmonics is larger than observed values for the Northern Hemisphere by a factor of about 4 (Northern Hemisphere zonal averages). These large amplitudes are due to the absence of ocean surface in the zonal averages. In the Northern Hemisphere, the land fraction is about 60% and in the Southern Hemisphere, the fraction is 80%. We will attempt to remedy this situation in Chapter 8 by introducing land–sea geography. Figure 6.7 shows the forcing for a circular orbit and the response superimposed in solid line contours. Note the lag of about a tenth of a year in the response. There is also a displacement poleward of the maximum response from the heating. Figure 6.8 indicates the response through time at some selected latitudes, this time including the effects of the eccentric orbit.

Image described by caption and surrounding text.

Figure 6.7 Illustration of the lag of zonal-average seasonal temperatures behind the forcing for the all-land planet. Shown is a contour plot (solid contours) of seasonal temperature response for the circular orbit insolation in dashed contours superimposed on the forcing to show the lag between heating and response. For a mixed-layer all-ocean planet where the response time is several years, the lag will be close to c06-math-113 radians or 0.25 year.

The response through time at some selected latitudes, including the effects of the eccentric orbit plotted for 37N, 24N, 11.5N, 5.7N, 0N, 5.7S, 11.5S, 24S, and 37S.

Figure 6.8 EBM solutions for the seasonal cycle of c06-math-114 for the all-land planet with present orbital parameters at selected latitudes: c06-math-1150.6 (36.9N/S), c06-math-1160.4 (23.6N/S), c06-math-1170.2 (11.5N/S), c06-math-1180.1 (5.7N/S) and 0 (equator).

6.4 Spread of Diffused Heat

Since EBMs make use of diffusion as a mechanism to transport heat poleward in the atmosphere/ocean system, it is useful to see how diffusion is related to random walk processes. Random walk means that the progress (root mean square average distance from the point of origination) of a passive scalar is the sum of a large number of steps. Let c06-math-119 be the random variable denoting the displacement after c06-math-120 steps. We can write

6.27 equation

The mean over an ensemble of such random walks is 0, because we assume the individual steps have mean 0, that is,

6.28 equation

The variance of c06-math-123 can be calculated if all the c06-math-124 are uncorrelated and have equal variance:

6.29 equation

The standard deviation of c06-math-126 is proportional to the square root of the number of steps.

Now consider the one-dimensional damped diffusion equation

6.30 equation

It is convenient to solve this equation with Fourier transforms. c06-math-128 can be represented as follows:

6.31 equation

We can write the second spatial partial derivative as follows:

6.32 equation

The inverse Fourier transform is

6.33 equation

We can insert the other terms.

6.34 equation

The integrand must vanish for every wave number.

6.35 equation

The solution to this first-order homogeneous linear equation for wave number c06-math-134 is

If the initial distribution is peaked at c06-math-136, that is, c06-math-137, then c06-math-138, a constant. The inverse Fourier transformation

6.37 equation

gives (from tables or MATHEMATIC A)

6.38 equation

where we have used the now familiar length scale c06-math-141 and c06-math-142 to make the notation more compact and to see the variables c06-math-143 and c06-math-144 proportional to their natural scales, c06-math-145 and c06-math-146. We see that except for the overall damping factor c06-math-147, the solution spreads like a bell-shaped curve with standard deviation

6.39 equation

The interpretation is that heat spreads symmetrically away from a point-concentrated initial anomaly a distance that is comparable to the damped diffusive length scale c06-math-149 in about one characteristic time c06-math-150. This is shown in Figure 6.9, wherein the spatial units are proportional to c06-math-151 and temporal units are proportional to c06-math-152.

Three dimensional plot with time, temperature, and distance on the axes of temporal evolution of a point pulse of heat at the origin at time equal to 0.

Figure 6.9 Temporal evolution of a point pulse of heat at the origin at time equal to 0. Horizontal distance in units of c06-math-153, temporal span in units of c06-math-154.

6.4.1 Evolution on a Plane

This section generalizes the treatment to two horizontal dimensions on the infinite c06-math-155c06-math-156 plane. We begin by including diffusion in two Cartesian dimensions:

6.40 equation

This time we use the two-dimensional Fourier transform pair (in the two spatial dimensions):

6.41 equation

where the two-dimensional vectors c06-math-159 and c06-math-160 have been introduced. Following the steps from the subSection 6.4, we find

6.42 equation

where c06-math-162, and the integration limits c06-math-163 have been suppressed. We proceed by using polar coordinates in the c06-math-164 plane. We write c06-math-165 where c06-math-166 is the angle between the vectors c06-math-167 and c06-math-168, and c06-math-169.

6.43 equation

We proceed by considering first the portion of the double integral into its c06-math-171 part (enclosed in parentheses): This angular integral is well known (it is a special case, c06-math-172, of Bessel's integral (see Whittacker and Watson, 1962 p. 362; or other books on mathematical physics)):

6.44 equation

where c06-math-174 is the Bessel function. The resulting integral can also be solved (e.g., MATHEMATICA), yielding the rather unsurprising answer:

6.45 equation

Note that in this compact form c06-math-176 is expressed proportional to the damped diffusion length scale c06-math-177 and c06-math-178 is always found proportional to the global timescale c06-math-179. We have found exactly what we found in one dimension: a delta function point expanding into a 2-D Gaussian shape, further expanding its disk width in a time c06-math-180 that is about c06-math-181 in radius. The volume under the Gaussian surface also diminishes by the factor c06-math-182 owing to the damping from infrared radiation to space.

6.5 Random Winds and Diffusion

Next consider the one-dimensional energy-balance equation in which heat is advected by a wind field that for simplicity has no c06-math-183-dependence.

6.46 equation

Once again, look at the component of the wave number c06-math-185

6.47 equation

This is a first-order linear equation and can be solved using the usual integrating factor:

6.48 equation

Let c06-math-188 be a stationary random function of time. The ensemble average of a random quantity is denoted by c06-math-189. Let the ensemble average vanish; stationarity demands that c06-math-190 with c06-math-191. This means that c06-math-192 will have mean zero and be stationary as well. Finally, we assume that c06-math-193 is a Gaussian random field. In other words, our wind field is similar to a random eddy field that carries heat one way or another depending on the vagaries of such a wind field. Note, however, that we have not permitted the wind to be a function of position. This makes our wind field a rather peculiar one: at a point in time it is everywhere pointed in the same direction, until the next instant, when it suddenly switches direction and magnitude the same everywhere. The case of the wind field that is variable in space as well as time is more difficult to solve. Rather than going into the complexities involved in that, we proceed with what we have.

Consider the series expansion of the second exponential factor, c06-math-194.

The first three terms are

6.49 equation

The second term vanishes because c06-math-196. In fact, all the odd powered terms vanish.a We turn to the third term after running the integrals from time c06-math-197 to c06-math-198 instead of 0 to c06-math-199. This symmetry helps in analyzing the result.

Image described by caption and surrounding text.

Figure 6.10 Geometry for a two-dimensional integral over a lag covariance function for a stationary process: c06-math-238. Since c06-math-239 only depends on c06-math-240, we can find the integral by summing the diagonal infinitesimal strip whose width is c06-math-241. The value of c06-math-242 in the lower-right corner is c06-math-243 and its value in the upper-left corner is c06-math-244. The quantity to be summed is the length of the diagonal strip (c06-math-245) times c06-math-246. The integral becomes c06-math-247. (Redrawn from original figures in Papoulis (1984).)

6.50 equation

The double integral can be computed by a transformation of variables from c06-math-201 to a one-dimensional integral. Referring to Figure 6.10, as the integrand only depends on c06-math-202, the integration is best done by summing slabs that are at a 45c06-math-203 angle (along which c06-math-204) to the horizontal. The length of the slab can be found from the hypotenuse of the right isosceles triangle whose lower edge has length c06-math-205. We can show this last by noting that the same side has length c06-math-206 where c06-math-207 is evaluated at the point the hypotenuse intersects the line c06-math-208. At that point, c06-math-209, or c06-math-210. Using c06-math-211 we finally arrive at the width of thelower side of the triangle to be c06-math-212. The length of the slab is c06-math-213. The width of the slab is c06-math-214; the two square roots cancel to leave us with

6.51 equation

where c06-math-216 is the autocorrelation function for the magnitude of the lag c06-math-217. Note that c06-math-218 and c06-math-219 is the variance of the velocity field. We specify that the integration time c06-math-220 is very long compared to the autocorrelation time c06-math-221 that we take to be c06-math-222 in Figure 6.11. This means that the kernel (nearly an isosceles triangle of base width 2c06-math-223 and height unity) of the integral is concentrated near c06-math-224 and the integral is c06-math-225, which is

Image described by caption and surrounding text.

Figure 6.11 Depiction of how a short autocorrelation time leads to an estimate of a double integral related to second moments.

(Redrawn from original figures in Papoulis (1984).)

6.52 equation

After replacing c06-math-227 by c06-math-228 to regain the original notation, we can now identify

6.53 equation

which gives us a handy formula for the diffusion coefficient in terms of the variance of the wind speed and its autocorrelation time. The last expression is reminiscent of the form of (6.36). Note that the units of c06-math-230 are (length)c06-math-231(time)c06-math-232, when the diffusion equation is set with the coefficient of c06-math-233 set to unity. When the approximations above are valid, we can think of diffusion equations as being the equations describing the ensemble averages of (damped) random walk equations. Basically, the approximation is that the winds are white noise (c06-math-234 days) compared to our averaging times. All timescales, including relaxation time of a column of air (c06-math-23530 days), and forcing functions such as the seasonal cycle (few months), must be long compared to the weather–noise timescale (few days). This suggests the following:

6.54 equation

When this approximation is valid, we can safely think of the EBCM equation as an equation for the ensemble average of c06-math-237. Note that there is a residual whose ensemble average is zero. Later, we will use this residual as a noise driver of climate fluctuations.

6.6 Numerical Methods

6.6.1 Explicit Finite Difference Method

We start with the energy-balance equation:

6.55 equation

where

6.57 equation

The coefficient in (6.51) was constructed by multiplying numerator and denominator by c06-math-251, then using c06-math-252 and c06-math-253. We start with a grid from c06-math-2541 to 1, divided into c06-math-255 equal segments. More details on the grid are presented in the following. The centered finite difference form for the derivative is

6.58 equation

The derivative of c06-math-257 is then

6.59 equation

where

6.60 equation

To step forward in time, we have the following algorithm (in the explicit finite difference case):

where the important parameter c06-math-261 is given by

6.62 equation

Next, we must make sure our solution enforces the Neumann boundary conditions, implying that no net heat flux enters the infinitesimal latitude circles surrounding the poles:

6.63 equation

The fact that the time-dependent version of the equation has a regular singular point at the pole poses a problem because as noted earlier, small numerical errors will lead to erroneous encroachment by the irregular solutions (c06-math-264). This does not turn out to be a serious problem as we will see. The most straightforward way of representing the equation in finite difference form is to break the interval c06-math-265 into c06-math-266 intervals, c06-math-267. Similarly, the time is discretized in steps of c06-math-268.

Deferring discussion of the boundary points, we see that the term on the left is the value of c06-math-269 at time c06-math-270, while all the terms on the right are to be evaluated at the previous time c06-math-271. This is very convenient because we presumably have knowledge of the field at time c06-math-272. For example, if we start at time c06-math-273, we know the initial profile of the field. This process allows us to evaluate it at c06-math-274 and so on. Such an algorithm is very appealing intuitively as it feels like we are solving the equation just as nature does it. Note that there is a problem at the end points for the reason that when c06-math-275, the RHS contains the value c06-math-276 and when c06-math-277, it contains the value c06-math-278. Hence, we must use the aforementioned algorithm only for c06-math-279. Had the boundary conditions been of the Dirichlet type where the values of the field are specified at the boundaries, we could merely specify the values of c06-math-280 and c06-math-281.

Enforcing the Neumann boundary conditions is a little tricky as the straightforward implementation of the thinking above leads to c06-math-282 and c06-math-283. We can get around this by using a centered difference of width c06-math-284 for these points. We introduce fictitious points just outside the range, c06-math-285, and c06-math-286. Then the boundary conditions read as follows:

6.64 equation

We compute the outside points using the explicit algorithm, but then force c06-math-288 to equal the lower outside value c06-math-289 and likewise c06-math-290 is forced to be c06-math-291.

The curves in Figure 6.12 show an example of implementation of this kind of algorithm for c06-math-292, where we have modified the definition of the parameter c06-math-293 to include c06-math-294, that is, c06-math-295, thereby allowing the stability parameter to be a function of latitude. The forcing is c06-math-296. The initial condition for the integration is c06-math-297C. There are 20 intervals in c06-math-298 from pole to pole, and the time step is c06-math-299 of the relaxation time c06-math-300. The figure shows a sequence of profiles after 10, 20, 40, 80, 160, 320, and 640 time steps. The last is nearly indistinguishable from the 320 time-step case. The solutions in this case are smooth and stable after integration to 3.20 relaxation times. Advancement to 3000 steps (30 relaxation times) indicates no change. We infer stability of the solution.

Image described by caption and surrounding text.

Figure 6.12 Sequence of the evolution toward steady state where the forcing is the usual c06-math-314. The initial condition is flat with c06-math-315C. The time step is 0.05, which is 1/20 of the relaxation time, c06-math-316. The individual graphs are at c06-math-317, and 640 steps. In this case, c06-math-318W mc06-math-319C)c06-math-320. An additional run (not shown) of length 3000 steps showed no change.

If we increase the time step by a factor of 10 to 0.1c06-math-304, we run into trouble. This case is shown in Figure 6.13 starting from the same initial condition. The stability parameter c06-math-305 increases to 3.20. In the integration, everything goes well until time step 400 in which a small ripple occurs at the equator. By time step 425, the instability is full blown and propagating from the equator toward the poles. The reason for this peculiar behavior is that the explicit method is unstable when the time steps are too large compared to the spatial increment. For the diffusion equation, the condition can be made quantitative: the solution will be unstable if c06-math-306. Since c06-math-307 is largest at the equator, the instability breaks out there first as indicated in the figure. The physical interpretation of the instability is that for diffusion or random walk processes, information propagates an rms distance c06-math-308. Turning this around, we have c06-math-309; if c06-math-310 in the numerical integration is larger than the time for propagation of information in the continuous exact solution, we can expect instability of the numerical procedure. A similar limitation occurs in the integration of wavelike (hyperbolic) equations with wave motion entering as the propagation mechanism (c06-math-311, c06-math-312 = wave speed).

Image described by caption and surrounding text.

Figure 6.13 The smooth line represents the approximate solution after 300 time steps where the approximate solution seems to have settled down to the correct answer. But if the integration continues to 425 steps, one encounters the spiky line indicating numerical instability. In this case, c06-math-301W mc06-math-302C)c06-math-303.

The finite difference algorithm as noted above in (6.56) is not the only one that is accurate with errors of order (c06-math-313). The next subsection considers some alternatives that have different numerical stability properties.

6.6.2 Semi-Implicit Method

An intermediate method is often used in practice. It is simply the weighted average of the explicit and implicit algorithms:

6.65 equation

with c06-math-322.

One has to gather the coefficients of c06-math-323 and place them onto the left-hand side. The coefficient matrix has to be inverted to obtain c06-math-324. The result is

c06-math-325

where c06-math-326 is a large matrix with c06-math-327 and c06-math-328 spatial indices. The matrix has to be inverted to find the next time-step values as a function latitude as indicated by the index c06-math-329: c06-math-330. Fortunately, the matrix c06-math-331 is usually sparse (most entries are zero), and very fast algorithms are available. A special case that is commonly used is for c06-math-332, which is called the Crank–Nicolson scheme. Stability properties can be found in books on numerical analysis, but experience shows that use of this method allows larger time steps before instability sets in. Semi-implicit algorithms are commonly used in numerical solution of general circulation models.

In the finite difference solution to more complicated models treating dynamics of the flows as well as radiation, and so on, it is found to be advantageous to treat some terms on the RHS of the equation with one semi-implicit weighting and other terms with different weightings. These are called splitting methods. Again, they are commonly used in numerical solutions of climate models.

6.7 Spectral Methods

6.7.1 Galerkin or Spectral Method

An alternative to the finite difference methods is to write an approximate solution to the field as a series of (usually) orthogonal functions

6.66 equation

where c06-math-334 is some finite cutoff to the series. A rather obvious choice for the c06-math-335 for this class of problems is the Legendre polynomials, c06-math-336. In the present case, this renders the problem trivial as the c06-math-337 are the eigenfunctions of c06-math-338. The problem becomes nontrivial if the diffusion coefficient depends on c06-math-339. Such an EBM may be written as

6.67 equation

with the usual Neumann boundary conditions at the poles. If c06-math-341 is inserted and each side is multiplied by c06-math-342 and integrated from pole to pole with respect to c06-math-343, we obtain

6.68 equation

where

6.69 equation

Now the problem has been reduced to a set of c06-math-346 first-order coupled ordinary differential equations for the time-dependent coefficients c06-math-347. The coupling matrix is symmetric, hence in this case we can even find its eigenvectors and conduct a stability analysis. From discussions earlier in this chapter, it is easy to relate the eigenvalues to relaxation times for the eigenmodes of the problem. The numerical integration will be unstable if the time constant for the highest eigenmode retained (c06-math-348th) is shorter than the time step employed in the time-stepping algorithm. In other words, the criterion encountered in the explicit scheme comes up again unless special precautions are taken.

Other basis sets are possible besides the Legendre polynomials. For example, one might try a Fourier series

6.70 equation

This choice is equivalent to the use of Chebyshev polynomials. This method is successful in the present one-dimensional case but requires considerable work and extra discussion when applied to the case of two horizontal dimensions. One advantage is that transformations from grid points to components c06-math-350 and vice-versa can be accomplished via fast Fourier transform.

6.7.2 Pseudospectral Method

A complication arises in the spectral method if one of the coefficients is space dependent, for example, c06-math-351. Then,

6.71 equation

with

6.72 equation

and

6.73 equation

where

6.74 equation

and the c06-math-356 are known as interaction coefficients. They can be tabulated or generated from recurrence relations. It is important to notice how many are necessary in a high-resolution simulation. For example, in a GCM simulation the truncation level might be 15 or 42 typically. This means there are 3375 or 74 088 coefficients that have to be stored in a lookup table. Some storage savings can be gained by noting that most of the interaction coefficients are 0. However, when the problem is elevated to two dimensions on the sphere, it becomes truly formidable. One way to get around it is to use a so-called pseudospectral method.

In the pseudospectral method, we transform back and forth between a grid point representation and the spectral representation. The grid point representation for the sphere involves use of the Gaussian quadrature method of numerically estimating integrals. An integral may be estimated by the sum

6.75 equation

where the c06-math-358 are weights associated with each term and the c06-math-359 are unequally spaced but specified points along the axis. The c06-math-360 and c06-math-361 are optimal for a given level c06-math-362. Gaussian integrationb of order c06-math-363 has the property that it integrates a c06-math-364 degree polynomial exactly. The abscissas c06-math-365 are the c06-math-366 zero of c06-math-367 and the weights c06-math-368 are c06-math-369. As an example, Figure 6.14 shows a plot of c06-math-370 (solid gray line) along with a plot of c06-math-371 (black line) with heavy points on the dashed line corresponding to the value of the weights at the roots c06-math-372 of c06-math-373. Next, the quantities c06-math-374 and c06-math-375 are calculated by

6.77 equation
Image described by caption and surrounding text.

Figure 6.14 Plot of c06-math-378 (solid gray curve) with zeros at c06-math-379. Also a plot of c06-math-380 (heavy black U-shaped curve) with vertical thin black lines corresponding to the value of the weights at the roots c06-math-381 of c06-math-382. The ordinate values of the intersections are the weights. The graph is only shown for c06-math-383, as all the functions are even.

The next time step may now be evaluated:

6.78 equation

where the ratio on the LHS stands for time advancement by some ODE algorithm such as Runge–Kutta. When the new set c06-math-385 are computed, we can reevaluate c06-math-386 by

The procedure delineated in (6.76)–(6.79) can be repeated as many times as necessary as long as the employed ODE algorithm is stable.

Studies have shown that the pseudo-spectral method pays over the interaction coefficient method when the number of stored coefficients exceeds a few tens of thousands. Since this is rather quickly exceeded in GCM calculations, the pseudospectral method is commonly used in numerical simulations.

6.8 Summary

This chapter has introduced time into the energy balance models. The first result is that for the linear zonally symmetric models, we find that if a climate is perturbed from equilibrium and then released, it relaxes back to its steady-state solution. The relaxation can be characterized as a sum of contributions from relaxation modes, each having an exponential decay with large scales having longer relaxation times than smaller ones. The zonally symmetric planet can also be solved for its seasonal cycle. It is remarkable that the solar insolation can be represented for many purposes as a sum of terms involving only the Legendre modes 0, 1, and 2. The largest part of the circular orbital insolation is controlled by Legendre mode of index 1, which is antisymmetric on the globe and driven by a single annual cycle sinusoid in time. The Legendre mode of index 2 also has a non-negligible contribution in the semiannual cycle representing the passage of the Sun over the equator twice per year. These simple facts allow the seasonal cycle to be solved for this configuration. We can readily study the major effects on insolation of changing orbital parameters. The amplitude of the annual cycle is much too large if the zonally symmetric planet is considered to be all-land and too smallif all-ocean. We postpone the study of land–sea distribution until Chapter 8.

Next in the chapter was a study of how a patch of heat deposited in a small area spreads laterally to larger circular areas in a time roughly that of the characteristic time of the global model, c06-math-388 to a radius of the characteristic length c06-math-389. The spread has an exponentially damped (time constant c06-math-390) Gaussian shape with standard deviation proportional to c06-math-391. It was also shown that if the diffusion term is replaced by a random-wind advection term, the ensemble result will be the same as diffusion. If the autocorrelation time of the velocity field is short compared to other times, such as c06-math-392, in the problem, then the ensemble average of this term is essentially the diffusion term. This is very close to the Brownian motion problem solved by Einstein in 1904.

The chapter concludes with a study of how 1-D EBMs can be solved by finite difference methods. In particular, an explicit problem is worked through in some detail with an illustration of how the procedure becomes unstable if the time step exceeds a certain threshold depending on the characteristic time and length scales of the EBM. Basically, the time step must be shorter than the spreading time across a grid box in the horizontal direction.

The appendix to the chapter provides some derivations of the orbital dependence of the insolation function.

6.8.1 Parameter Count

We have introduced the time dimension, which means we must now have an effective heat capacity that cannot be calculated from first principles. For the all-land planet and at frequencies around 1/year, our guess is to take it to be 1/2 the heat capacity of the air column's mass at constant pressure. This neglects the heat capacity of soil. We are also neglecting the topography, where atop mountains, the effective heat capacity might be less. For an all-ocean planet we take the effective heat capacity to be that of the ocean's mixed layer. It depends on position, especially latitude, but we take it to be constant. So for the latitude-only model we have introduced one new parameter, but we have learned a lot about how the seasonal cycle works for a uniform planet.

Notes for Further Reading

Papoulis (1984) is an excellent book on random processes and spectral analysis written mainly for electrical engineers. It is rather compact, but comprehensive. Numerical methods for partial differential equations are covered in Ames (1992) and Fletcher (1991). Multigrid methods for solving partial differential equations are covered in Hackbusch (1980) and Briggs et al. (2000) and applied to EBMs by Bowman and Huang (1991), Huang and Bowman (1992), and Stevens and North (1996). Mars has no ocean, so a constant heat capacity can be used on the (approximately) homogeneous planet. A one-dimensional model with a zonally symmetric seasonal cycle was used successfully by James and North (1982). Much more can be found about the Martian climate as well as the other planets in Ingersoll's recent book on planetary climates Ingersoll (2015). Planetary climates are covered at a higher level of physics by Pierrehumbert (2011).

Exercises

  1. 6.1 Given the time-dependent, one-dimensional energy balance model
    equation

    where c06-math-393, and coalbedo c06-math-394 are constants, derive the basic timescale and spatial scale of the model by nondimensionalizing the model.

  2. 6.2 Consider the time-dependent, one-dimensional model of the previous exercise, where the parameters, c06-math-395, and c06-math-396 are constants. Let the solution of the energy balance model be denoted as follows: c06-math-397. An external forcing is applied at time c06-math-398 so that the temperature field is perturbed into the form c06-math-399. The external forcing disappears instantaneously (delta function in time). (a) Derive the governing equation for c06-math-400. (b) Solve the governing equation to obtain the solution for c06-math-401.
  3. 6.3 Consider next the same one-dimensional model as before with the same constant parameters. Let the solar distribution function be decomposed as follows:
    equation

    where Re(c06-math-402) denotes the real part of its argument. Solve the energy balance model for c06-math-403.

  4. 6.4 Show that the solution of the one-dimensional damped diffusion equation
    equation

    with the initial condition c06-math-404 is given by

    equation

    where c06-math-405 and c06-math-406.

  5. 6.5 A semi-implicit algorithm for a time-dependent, one-dimensional energy balance model is given by

    c06-math-407

    where

    equation

    and c06-math-408 and c06-math-409 denote a point in space and time, respectively. This equation can be rewritten as

    equation

    Determine the matrices c06-math-410 and c06-math-411 explicitly.

  6. 6.6 Consider a time-dependent, one-dimensional energy balance model in the form
    equation

    where c06-math-412 represents the constant heat capacity and c06-math-413. Set up the solution procedure using the spectral method. Use the Legendre polynomials as basis functions.

  7. 6.7 In Chapter 5, the insolation distribution function was determined by using spherical geometry. Consider the situation depicted in Figure 6.15. (a) Determine the normal vector at point c06-math-414 and at point c06-math-415 in terms of the longitude, latitude, and declination angle, and unit vectors i , j, and k. Then, calculate the cosine of the angle between the two vectors. (b) Determine the range of longitude for which the Sun is visible. (c) Based on your answers for (a) and (b), determine the total amount of insolation received at the surface when the solar constant is c06-math-416.
    Image described by caption and surrounding text.

    Figure 6.15 Diagram for Exercise 6.7.

6.9 Appendix to Chapter 6: Solar Heating Distribution

This chapter made use of the solar heating (insolation) function c06-math-417, which is the amount of radiant energy per unit time per unit surface area averaged through the day reaching the top of the atmosphere. In this appendix, we present a short derivation of this function, as its development into Legendre functions in latitude and sinusoids in time helps to understand the excitation of these modes in the response field under different orbital conditions. The appendix is organized as follows. First the derivation of c06-math-418 is given, followed by a discussion of how the orbital elements enter the mode amplitudes. Figure 6.16 shows an octant of the Earth's surface with a given point c06-math-419 singled out for our attention. The North Pole lies in the c06-math-420 direction, and hence the point c06-math-421 will rotate uniformly around the c06-math-422-axis in the course of a day. As the day passes, c06-math-423 will vary linearly from 0 to c06-math-424. At a given time of day, c06-math-425, the amount of solar power per unit area deposited at c06-math-426 is given by the solar constant c06-math-427 times the cosine of the angle between zenith direction c06-math-428 and the line joining the Sun and the Earth c06-math-429. In other words, c06-math-430. We define the solar vector c06-math-431 to lie in the c06-math-432 plane. Using the Cartesian unit vectors c06-math-433, and c06-math-434, we may write

6.80 equation
6.81 equation

where c06-math-436 is the angle the solar vector makes with the equatorial plane (c06-math-437c06-math-438 plane). Clearly, c06-math-439 depends on the time of the year. We readily compute

6.82 equation
Image described by caption and surrounding text.

Figure 6.16 An octant of the Earth's surface showing a unit vector toward the Sun c06-math-441, lying in the c06-math-442 plane. A given point on the Earth's surface is designated c06-math-444. The North Pole is along the c06-math-445-axis. The local time of day at point c06-math-446 is proportional to c06-math-447. The declination c06-math-448 is the angle the Sun makes with a perpendicular to the equatorial plane (the c06-math-449c06-math-450 plane here) for a given day of the year, c06-math-451. The declination depends on the tilt of the Earth's axis with respect to a perpendicular to the plane of the ecliptic (the obliquity) and the time of the year.

Image described by caption and surrounding text.

Figure 6.17 Half length of daylight in radians as a function of time of the year (c06-math-452) and the polar angle c06-math-453 for obliquity c06-math-454 and eccentricity c06-math-455. When the half day c06-math-456, it indicates perpetual daylight. The obliquity is chosen to be 35c06-math-457 (as opposed to its present value of 23.47c06-math-458) for illustration.

Image described by caption and surrounding text.

Figure 6.18 Graphic of the seasonal cycle of heating energy flux at the top of the atmosphere for a circular orbit and obliquity c06-math-459 as a function. The function is normalized by the total solar irradiance c06-math-460. (a) The three mode (c06-math-461) insolation function. (b) The exact solution as described in Appendix A. A larger value of obliquity (c06-math-462) is used to illustrate the polar day and night characteristics in the figure.

To obtain the diurnal average of the solar power at c06-math-463 per unit area, we must integrate c06-math-464 through the daylight hours (the range of c06-math-465 for which c06-math-466) and divide by the length of the whole day (c06-math-467). Dawn and dusk can be defined as c06-math-468, the roots of c06-math-469.

The half day is shown in Figure 6.17. The formula can now be written for the solar power per unit area averaged through the day,

where c06-math-472 differs slightly from the total solar irradiance (because it is not annualized). Because of the elliptical orbit,

6.85 equation

where c06-math-474 is the Earth–Sun distance at the given time of year and c06-math-475 is the annual average of c06-math-476. Figure 6.18 shows the seasonal cycle of insolation function for a circular orbit.

6.9.1 The Elliptical Orbit of the Earth

The Sun lies at the focus of an ellipse that constitutes the Earth's trajectory over the year (Figure 6.19). Using the Sun as the origin of a polar coordinate system with c06-math-477 as polar angle in the ecliptic plane (celestial longitude), we can write

6.86 equation

where c06-math-479 is the eccentricity (presently c06-math-480),

6.87 equation

and c06-math-482 is very nearly the average of c06-math-483 (to fourth order in powers of c06-math-484). The value of c06-math-485 determines the time of year of the closest approach of the Earth to the Sun (perihelion). If we arbitrarily choose that c06-math-486 at winter solstice (c06-math-487December 22 in the present epoch), then at present c06-math-488, as perihelion occurs at present within a few weeks of (Northern Hemisphere) winter solstice. However, c06-math-489 increases linearly through c06-math-490 in a time of 22 000 years, leading to a passing of perihelion through the seasons with a period of 22 000 years.

In what follows, we can take c06-math-491 to be fixed and consider the time dependence of c06-math-492. If the Earth's orbit were circular, c06-math-493 would be linear with time:

6.88 equation

Since the Earth's orbit is really elliptical, we must use the conservation of angular momentum to compute c06-math-495. This is expressed as c06-math-496. After some manipulation, it can be shown that, to the first order in powers of c06-math-497,

6.89 equation
Image described by caption and surrounding text.

Figure 6.19 Diagram of the Earth's orbit illustrating the orientation of various unit vectors. The vector c06-math-499 points out along the North polar axis. The vector c06-math-500 is the position vector of a given point on the Earth's surface. The declination on a given day is the angle between c06-math-501 and c06-math-502, that is, c06-math-503. The unit vector c06-math-504 is perpendicular to the plane of the Earth's orbit (the ecliptic plane). The unit vector c06-math-505 points from the Earth's center to the Sun. The angle c06-math-506 is the celestial longitude.

6.9.2 Relation Between Declination and Obliquity

The obliquity or tilt of the Earth's orbit is the angle between c06-math-507 and c06-math-508 or c06-math-509. At present, this angle is about c06-math-510. The declination on a given day is the angle between c06-math-511 and c06-math-512, that is, c06-math-513. With some straightforward geometry, we can show that

6.90 equation

6.9.3 Expansion of c06-math-515

Consider now the expansion of c06-math-516 into a series of c06-math-517 and c06-math-518, the latter being a Fourier series as the function is strictly periodic in c06-math-519 (units of years here).

6.91 equation

In principle, the coefficients c06-math-521 can be computed numerically from (6.84) and (6.83); however, the main seasonal driving term can be computed analytically:

6.92 equation
6.93 equation

Furthermore, it can be shown that

6.94 equation

Table 6.1 Fourier–Legendre coefficients for the present distribution of incident solar radiation c06-math-524

c06-math-525 c06-math-526 c06-math-527 c06-math-528
0 0 1.0001 0.0000
1 0.0327 0.0067
2 0.0006 0.0003
1 0 0.0000 0.0000
1 −0.7974 −0.0054
2 −0.0261 −0.0052
2 0 −0.4760 0.0000
1 −0.0180 −0.0026
2 0.1486 −0.0022
4 0 −0.0444 0.0000
1 −0.0029 0.0000
2 0.0909 −0.0012

a The coefficients are zero for odd values of c06-math-529 greater than one. c06-math-530 corresponds to the Northern Hemisphere winter solstice.

Table 6.1 presents the first few coefficients for the present orbital parameters. Truncating the series at c06-math-531 and c06-math-532 is an excellent approximation for many purposes. Note that the c06-math-533 is due to the eccentric orbit. In fact, c06-math-534, where the present eccentricity is about 0.016. More numerical values of orbital changes can be found in North and Coakley (1979).

..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset
18.219.86.155