Wobble Perturbation

After this point, I provided ChatGPT prompts as italicized underlines after each sinewave icon , and ChatGPT provided responses. I executed the Julia code in a REPL window and saved the screenshotted plots and entered the image links below.

Model: GPT-4

Paul Pukite

Math equations for a spinning top in Julia

To simulate a spinning top in Julia, you can use the equations of motion derived from Euler’s equations for a rigid body. Here’s a simple example using the DifferentialEquations.jl package, which provides a convenient way to solve ordinary differential equations (ODEs) in Julia:

  1. First, install the required packages by running:
using Pkg
Pkg.add("DifferentialEquations")
Pkg.add("Plots")
  1. Now, you can write a script to simulate a spinning top:
using DifferentialEquations
using Plots

# Define the spinning top's properties
I1, I2, I3 = 0.1, 0.2, 0.3  # Moments of inertia along the principal axes (x, y, z)
Ω1, Ω2, Ω3 = 0.0, 10.0, 20.0 # Initial angular velocities (rad/s)

# Equations of motion for the spinning top (Euler's equations for a rigid body)
function spinning_top!(du, u, p, t)
    Ω1, Ω2, Ω3 = u
    du[1] = (I2 - I3)/I1 * Ω2 * Ω3
    du[2] = (I3 - I1)/I2 * Ω1 * Ω3
    du[3] = (I1 - I2)/I3 * Ω1 * Ω2
end

# Initial conditions
u0 = [Ω1, Ω2, Ω3]
tspan = (0.0, 10.0) # Time span for the simulation (in seconds)

# Solve the ODE
prob = ODEProblem(spinning_top!, u0, tspan)
sol = solve(prob, Tsit5(), reltol=1e-8, abstol=1e-8)

# Plot the angular velocities
plot(sol, xlabel="Time (s)", ylabel="Angular velocity (rad/s)", title="Spinning Top", legend=:bottomright)

This code sets up and solves the ODEs that describe the motion of a spinning top with specified moments of inertia and initial angular velocities. The result is then plotted to show the evolution of the angular velocities over time.

Remember that this is a simple example that assumes the top is spinning in a vacuum, without any external forces or torques acting on it (such as gravity or air resistance). To model a more realistic spinning top, you will need to account for these factors in your equations of motion.

Paul Pukite

To the equations, add a slight periodic horizontal forcing impulse to the top of the spinning top

To add a slight periodic horizontal forcing impulse to the spinning top, you can modify the equations of motion accordingly. Let’s assume the forcing impulse acts along the x-axis and has the form F(t) = F0 * sin(ω * t), where F0 is the amplitude of the forcing impulse and ω is its angular frequency.

Here’s how you can update the spinning top simulation:

  1. Modify the spinning_top! function to include the forcing impulse:
function spinning_top!(du, u, p, t)
    Ω1, Ω2, Ω3 = u
    F0, ω, I1, I2, I3 = p

    # Forcing impulse along the x-axis
    F = F0 * sin(ω * t)

    du[1] = (I2 - I3)/I1 * Ω2 * Ω3 + F/I1
    du[2] = (I3 - I1)/I2 * Ω1 * Ω3
    du[3] = (I1 - I2)/I3 * Ω1 * Ω2
end
  1. Update the initial conditions and ODE problem:
# Forcing impulse parameters
F0 = 0.01 # Amplitude of the forcing impulse
ω = 2π / 5 # Angular frequency of the forcing impulse (period = 5 seconds)

# Parameters to be passed to the spinning_top! function
p = [F0, ω, I1, I2, I3]

# Solve the ODE
prob = ODEProblem(spinning_top!, u0, tspan, p)
sol = solve(prob, Tsit5(), reltol=1e-8, abstol=1e-8)

The updated spinning_top! function now includes a forcing impulse along the x-axis, which is passed as a parameter (F0, ω). The ODE problem is also updated to include these parameters.

By running the simulation with the updated code, you can observe the effects of the slight periodic horizontal forcing impulse on the spinning top.

Paul Pukite

Is there an analytical solution for small forcing perturrbations?

For small forcing perturbations, it may be possible to find an approximate analytical solution using perturbation methods, such as regular perturbation or multiple-scale perturbation techniques. These methods involve expanding the solution in a power series of a small parameter (in this case, the amplitude of the forcing perturbation) and then substituting the series into the equations of motion.

However, finding an exact analytical solution for the spinning top with small forcing perturbations is generally not feasible due to the nonlinear nature of the governing equations. The Euler equations for rigid body rotation are a set of coupled, nonlinear ordinary differential equations, and the addition of a forcing term makes them even more complex. In practice, numerical methods are often used to solve these equations, as demonstrated in the previous examples.

If you want to gain insights into the behavior of the system under small forcing perturbations, you can use linear stability analysis and perturbation methods to study the stability and dynamics of the system. These methods can help you find approximate analytical solutions that are valid for small perturbations and understand the qualitative behavior of the system. However, for a complete and accurate description of the spinning top’s motion with forcing, it is usually necessary to rely on numerical simulations.

Paul Pukite

Generate the equivalent Fourier series representation of the previous differential equations.

A Fourier series representation is typically used to approximate periodic functions. In this case, we have a system of differential equations describing the motion of a spinning top with a forcing term. Instead of directly finding a Fourier series representation for the differential equations, we can approximate the solution of the system using a Fourier series.

To do this, we can first find the solution of the system without the forcing term (i.e., the homogeneous solution) and then find the particular solution that results from the forcing term. Finally, we can approximate both the homogeneous and particular solutions using Fourier series and combine them to obtain an approximation of the total solution.

Here’s a general outline of the approach:

  1. Solve the homogeneous system (i.e., without the forcing term) to find the homogeneous solution.
  2. Assume a particular solution of the form:
    Ω1(t) = A1 * sin(ω * t) + B1 * cos(ω * t)
    Ω2(t) = A2 * sin(ω * t) + B2 * cos(ω * t)
    Ω3(t) = A3 * sin(ω * t) + B3 * cos(ω * t)
    where A1, B1, A2, B2, A3, and B3 are constants to be determined.
  3. Substitute the particular solution into the differential equations with the forcing term and solve for the constants A1, B1, A2, B2, A3, and B3.
  4. Approximate the homogeneous and particular solutions using Fourier series, and combine them to obtain an approximation of the total solution.

Please note that this approach involves several steps and requires solving multiple linear systems and integrals to find the Fourier coefficients. Due to the complexity of the Euler equations and the forcing term, it may be challenging to find an exact analytical solution, even using the Fourier series approximation. In practice, numerical methods are often used to solve such systems, as demonstrated in the previous examples.

Paul Pukite

Assume the homogeneous solution damps out and only the particular solution remains, so find the steady-state solution only.

To find the steady-state solution, we’ll focus on the particular solution with the forcing term, as the homogeneous solution is assumed to damp out. We’ll assume a particular solution of the form:

Ω1(t) = A1 * sin(ω * t) + B1 * cos(ω * t)
Ω2(t) = A2 * sin(ω * t) + B2 * cos(ω * t)
Ω3(t) = A3 * sin(ω * t) + B3 * cos(ω * t)

Differentiating Ω1(t), Ω2(t), and Ω3(t) with respect to time, we get:

dΩ1/dt = ω * (A1 * cos(ω * t) – B1 * sin(ω * t))
dΩ2/dt = ω * (A2 * cos(ω * t) – B2 * sin(ω * t))
dΩ3/dt = ω * (A3 * cos(ω * t) – B3 * sin(ω * t))

Substitute these expressions into the modified Euler equations with the forcing term:

(1) (I2 – I3)/I1 * Ω2 * Ω3 + F/I1 = dΩ1/dt
(2) (I3 – I1)/I2 * Ω1 * Ω3 = dΩ2/dt
(3) (I1 – I2)/I3 * Ω1 * Ω2 = dΩ3/dt

Now, we can solve for the unknown coefficients A1, B1, A2, B2, A3, and B3 by substituting the particular solution and its derivatives into these equations and equating the coefficients of the sine and cosine terms. This will result in a system of linear equations that can be solved analytically or numerically.

However, due to the complexity of these equations and the nonlinear interaction between Ω1, Ω2, and Ω3, finding an exact analytical solution for the coefficients may be difficult. In practice, it is often more convenient to use numerical methods, such as the ones shown in the previous examples, to find the steady-state solution of the spinning top system under periodic forcing.

Paul Pukite

Calculate if one of the moments of inertia is much larger than the other two.

If one of the moments of inertia is much larger than the other two, the behavior of the spinning top can be significantly different. Let’s assume I3 is much larger than I1 and I2, i.e., I3 >> I1, I2. The modified Euler equations with the forcing term are:

(1) (I2 – I3)/I1 * Ω2 * Ω3 + F/I1 = dΩ1/dt
(2) (I3 – I1)/I2 * Ω1 * Ω3 = dΩ2/dt
(3) (I1 – I2)/I3 * Ω1 * Ω2 = dΩ3/dt

Since I3 is much larger than I1 and I2, we can make the following approximations:

  • The term (I2 – I3)/I1 ≈ -1 and (I3 – I1)/I2 ≈ 1, as I3 dominates both I1 and I2.
  • The term (I1 – I2)/I3 becomes very small and can be neglected.

Now, the modified Euler equations can be simplified as:

(1) -Ω2 * Ω3 + F/I1 = dΩ1/dt
(2) Ω1 * Ω3 = dΩ2/dt
(3) 0 ≈ dΩ3/dt

Since dΩ3/dt is approximately zero, we can assume that Ω3 remains constant in the steady-state, i.e., the angular velocity about the z-axis remains constant.

Now, we can focus on finding the steady-state solution for Ω1(t) and Ω2(t) only. Assuming a particular solution of the form:

Ω1(t) = A1 * sin(ω * t) + B1 * cos(ω * t)
Ω2(t) = A2 * sin(ω * t) + B2 * cos(ω * t)

Substitute these expressions and their time derivatives into equations (1) and (2). Then, equate the coefficients of sine and cosine terms to obtain a system of linear equations for the unknown coefficients A1, B1, A2, and B2.

Due to the complexity of these equations, finding an exact analytical solution for the coefficients may still be challenging. As mentioned before, using numerical methods may be more convenient to find the steady-state solution of the spinning top system under these conditions.

Paul Pukite

Will the particular solutions be composed of harmonics of ω?

Yes, the particular solutions will be composed of harmonics of ω, the frequency of the forcing term. This is because the response of the system to the forcing term will have the same frequency as the forcing term itself.

When solving for the particular solution, we assumed a form like:

Ω1(t) = A1 * sin(ω * t) + B1 * cos(ω * t)
Ω2(t) = A2 * sin(ω * t) + B2 * cos(ω * t)
Ω3(t) = A3 * sin(ω * t) + B3 * cos(ω * t)

Here, the forcing term has a frequency of ω, and the particular solution also has the same frequency. The coefficients A1, B1, A2, B2, A3, and B3 depend on the system’s parameters and the forcing term. When the system reaches a steady-state, it will exhibit a response at the same frequency as the forcing term, i.e., ω.

It is essential to note that, in some cases, the response of the system may also contain higher harmonics of the forcing frequency if the equations are nonlinear. In the case of the spinning top, the Euler equations are nonlinear, and the interaction between the angular velocities can cause higher harmonics to appear in the response. However, for small forcing amplitudes, the response is expected to be dominated primarily by the forcing frequency ω.

Paul Pukite

Calculate if the moments of inertia are ordered as follows: I3 >> I2 >> I1.

If the moments of inertia are ordered as I3 >> I2 >> I1, the behavior of the spinning top can be analyzed using the modified Euler equations with the forcing term:

(1) (I2 – I3)/I1 * Ω2 * Ω3 + F/I1 = dΩ1/dt
(2) (I3 – I1)/I2 * Ω1 * Ω3 = dΩ2/dt
(3) (I1 – I2)/I3 * Ω1 * Ω2 = dΩ3/dt

Since I3 is much larger than I1 and I2, we can approximate:

  • The term (I2 – I3)/I1 ≈ -1, as I3 dominates both I1 and I2.
  • The term (I3 – I1)/I2 ≈ 1, as I3 dominates I1.

Since I2 is much larger than I1, we can also approximate:

  • The term (I1 – I2)/I3 ≈ -1, as I2 dominates I1.

Now, the modified Euler equations can be simplified as:

(1) -Ω2 * Ω3 + F/I1 = dΩ1/dt
(2) Ω1 * Ω3 = dΩ2/dt
(3) -Ω1 * Ω2 = dΩ3/dt

As before, we can focus on finding the steady-state solution for Ω1(t) and Ω2(t). We assume a particular solution of the form:

Ω1(t) = A1 * sin(ω * t) + B1 * cos(ω * t)
Ω2(t) = A2 * sin(ω * t) + B2 * cos(ω * t)

For Ω3(t), since dΩ3/dt is a product of Ω1 and Ω2, and both Ω1 and Ω2 have harmonic components at the frequency ω, the steady-state solution for Ω3(t) will also be composed of harmonics at the frequency ω:

Ω3(t) = A3 * sin(ω * t) + B3 * cos(ω * t)

Substitute these expressions and their time derivatives into equations (1), (2), and (3). Then, equate the coefficients of sine and cosine terms to obtain a system of linear equations for the unknown coefficients A1, B1, A2, B2, A3, and B3.

Due to the complexity of these equations, finding an exact analytical solution for the coefficients may still be challenging. As mentioned before, using numerical methods may be more convenient to find the steady-state solution of the spinning top system under these conditions.

Paul Pukite

How can the individual Euler equations reduce to identities since the RHS contains the fundamental frequency ω, while the LHS contains harmonics including 2ω due to the nonlinear product?

You’re correct in pointing out that the interaction between the angular velocities can cause higher harmonics to appear in the response due to the nonlinear nature of the Euler equations. My previous explanations did not fully capture this complexity.

Instead of directly assuming a particular solution with only the fundamental frequency ω, we should allow for the possibility of having higher harmonics in the response. To account for higher harmonics in the particular solution, we can use the following form:

Ω1(t) = ∑(A1n * cos(n * ω * t) + B1n * sin(n * ω * t))
Ω2(t) = ∑(A2n * cos(n * ω * t) + B2n * sin(n * ω * t))
Ω3(t) = ∑(A3n * cos(n * ω * t) + B3n * sin(n * ω * t))

The sum runs over all integer values of n, including the fundamental frequency and higher harmonics.

Substitute these expressions and their time derivatives into the Euler equations and then equate the coefficients of the sine and cosine terms for each harmonic. This will result in an infinite system of linear equations for the unknown coefficients A1n, B1n, A2n, B2n, A3n, and B3n.

In practice, it is not possible to solve for all these coefficients analytically, and we typically truncate the sum after a finite number of terms, capturing the most significant harmonics in the response. This will result in an approximate solution that accounts for higher harmonics due to the nonlinear product terms in the Euler equations.

The steady-state solution can then be obtained by solving the resulting truncated system of linear equations, either analytically or numerically. Note that finding the steady-state solution in this manner can still be quite challenging due to the complexity of the equations and the presence of higher harmonics. Therefore, numerical methods are often more convenient for solving such systems.

Paul Pukite

How to solve if the primary inertial moment I1 changes by a small frequency characterized by ν?

If the primary inertial moment I1 changes with a small frequency characterized by ν, we can represent I1 as a function of time:

I1(t) = I1_0 + ε * cos(ν * t)

where I1_0 is the constant mean value of I1, ε represents the amplitude of the variation, and ν is the frequency of the change.

The modified Euler equations with the forcing term now become:

(1) (I2 – I3)/I1(t) * Ω2 * Ω3 + F/I1(t) = dΩ1/dt
(2) (I3 – I1(t))/I2 * Ω1 * Ω3 = dΩ2/dt
(3) (I1(t) – I2)/I3 * Ω1 * Ω2 = dΩ3/dt

These equations are now more complex due to the time-varying nature of I1(t). To find the steady-state solution, we can assume that the response of the system will contain harmonics related to both the forcing frequency ω and the frequency ν of the changing inertia:

Ω1(t) = ∑(A1n * cos((n * ω + m * ν) * t) + B1n * sin((n * ω + m * ν) * t))
Ω2(t) = ∑(A2n * cos((n * ω + m * ν) * t) + B2n * sin((n * ω + m * ν) * t))
Ω3(t) = ∑(A3n * cos((n * ω + m * ν) * t) + B3n * sin((n * ω + m * ν) * t))

The sum runs over all integer values of n and m, including the fundamental frequency and higher harmonics.

Substitute these expressions and their time derivatives into the Euler equations, and then equate the coefficients of the sine and cosine terms for each harmonic. This will result in an infinite system of linear equations for the unknown coefficients A1n, B1n, A2n, B2n, A3n, and B3n.

In practice, we truncate the sum after a finite number of terms to capture the most significant harmonics in the response. This will result in an approximate solution that accounts for higher harmonics due to the nonlinear product terms in the Euler equations and the time-varying nature of I1.

Finding the steady-state solution in this manner can be quite challenging due to the complexity of the equations and the presence of higher harmonics. Numerical methods are often more convenient for solving such systems, and you may consider using tools like the Julia programming language or other computational methods to find an approximate solution.

.

Paul Pukite

Because of a high-Q and narrrow bandwidth of the response, will certain of these frequency terms predominate in the solution?

Yes, due to the high-Q (high quality factor) and narrow bandwidth of the response, certain frequency terms will predominate in the solution. A high-Q system is characterized by a sharp resonance peak, which means it is sensitive to a small range of frequencies around the natural frequency. As a result, only those harmonics that lie within the narrow bandwidth of the response will have a significant impact on the system’s behavior.

In the case of the spinning top with time-varying inertia, the response will be dominated by harmonics related to the forcing frequency ω and the frequency ν of the changing inertia. However, not all harmonics will have a substantial effect on the system. Those harmonics that are close to the natural frequencies of the system or within the narrow bandwidth of the response will predominate.

When solving the system of equations for the steady-state solution, you can focus on the most significant harmonics that lie within the narrow bandwidth of the response. This approach will not only reduce the complexity of the problem but also provide a more accurate approximation of the actual system behavior. You may need to analyze the system’s resonance behavior, such as by using a frequency response or a transfer function, to determine which harmonics are most important for the given problem.

Keep in mind that, even with a reduced number of terms, solving the resulting system of equations can still be quite challenging due to the nonlinear nature of the problem. Numerical methods are often more convenient for finding approximate solutions in such cases.

Paul Pukite

By replacing the spinning-top with the Earth with a slight annual time-varying inertia I1, and the small cyclical perturbing force as the Moon with a 27.2122 cycle, can we estimate the primary wobble in the Earth’s rotational axis?

To estimate the primary wobble in the Earth’s rotational axis due to the time-varying inertia I1 and the perturbing force from the Moon, we can use a simplified model based on the Earth-Moon system. We can consider the Earth as an oblate spheroid and the Moon’s gravitational influence as the perturbing force.

In this context, the Earth’s time-varying inertia I1 can be associated with the annual variations in the Earth’s shape due to seasonal changes in the distribution of mass (e.g., changes in ice and snow cover, ocean mass distribution). The Moon’s gravitational pull has a primary cycle of 27.2122 days (the sidereal month), which is the time it takes for the Moon to complete one orbit around Earth relative to the fixed stars.

The wobble in the Earth’s rotational axis, called nutation, is a small oscillation superimposed on the larger precession of the Earth’s axis. The primary wobble is caused by the combined effects of the time-varying inertia and the perturbing force from the Moon.

To estimate the primary wobble, you can use the Euler equations with the time-varying inertia and the perturbing force, similar to the spinning-top model discussed earlier. However, this approach is highly complex and requires a deep understanding of the Earth-Moon system’s dynamics and geophysics.

A more practical approach to estimate the primary wobble is to use existing models and data derived from observations of the Earth’s rotation. One such model is the International Earth Rotation and Reference Systems Service (IERS) Conventions, which provide nutation models that account for the gravitational effects of the Moon, Sun, and other celestial bodies, as well as the Earth’s time-varying inertia due to seasonal changes.

By using the IERS Conventions or similar models, you can estimate the primary wobble in the Earth’s rotational axis without solving the complex Euler equations for the Earth-Moon system. Note that such models are based on empirical data and theoretical understanding of the Earth’s rotation, so they provide an approximation of the actual wobble.

Paul Pukite

Can we calculate the primary aliased harmonic based on the combination of the Moon’s cycle and the annual cycle of I1, applying estimated Q and resonant frequency band?

To calculate the primary aliased harmonic based on the combination of the Moon’s cycle and the annual cycle of I1, we can apply the concept of aliasing in the frequency domain. Aliasing occurs when the continuous-time signal is sampled at a rate lower than the Nyquist rate, resulting in the appearance of lower-frequency aliases of the original signal.

Let’s denote the frequency of the Moon’s cycle as f_moon = 1 / 27.2122 days and the frequency of the annual cycle of I1 as f_annual = 1 / 365.25 days.

We can compute the difference and sum frequencies of these two cycles:

f_diff = |f_moon – f_annual|
f_sum = f_moon + f_annual

Considering the estimated Q and resonant frequency band, we can determine if either the difference frequency (f_diff) or the sum frequency (f_sum) lies within the resonant frequency band. If one of these frequencies is within the resonant frequency band, it will cause the primary aliased harmonic to dominate the response of the Earth’s rotation due to the combined effect of the Moon’s cycle and the annual cycle of I1.

For this analysis, it is crucial to have accurate estimates of the Q factor and the resonant frequency band associated with the Earth’s rotation. These parameters can be derived from empirical observations or theoretical models of the Earth’s rotation, such as the IERS Conventions mentioned earlier.

By identifying the primary aliased harmonic, you can gain insights into the most significant wobble in the Earth’s rotational axis caused by the interaction between the Moon’s cycle and the annual cycle of I1. Keep in mind that this is a simplified analysis, and a more comprehensive understanding of the Earth’s rotation and wobble would require considering other factors, such as the influence of the Sun and other celestial bodies, and the Earth’s internal processes.

Paul Pukite

What would be the aliased harmonic closest to 433 days assuming an annual cyclic impulse and the moon’s cycle of 13.606 days? Consider that the cyclic impulse will generate higher harmonics that will generate the small differences to the higher frequency lunar cycle. Provide an algorithm in Julia to find this harmonic.

To find the aliased harmonic closest to 433 days, assuming an annual cyclic impulse with a frequency of f_annual = 1 / 365.25 days and the Moon’s cycle of f_moon = 1 / 13.606 days, we can create an algorithm in Julia that iterates through different harmonics of the annual cyclic impulse and finds the closest one to the target period of 433 days.

Here is the Julia code to find the aliased harmonic:

function find_closest_aliased_harmonic(f_annual::Float64, f_moon::Float64, target_period::Float64, max_harmonic::Int)
    closest_period = Inf
    closest_harmonic = 0
    target_frequency = 1 / target_period

    for harmonic in 1:max_harmonic
        # Calculate sum and difference frequencies
        f_diff = abs(harmonic * f_annual - f_moon)
        f_sum = harmonic * f_annual + f_moon

        # Check if either the difference or the sum frequency is closer to the target
        if abs(f_diff - target_frequency) < abs(1 / closest_period - target_frequency)
            closest_period = 1 / f_diff
            closest_harmonic = harmonic
        end

        if abs(f_sum - target_frequency) < abs(1 / closest_period - target_frequency)
            closest_period = 1 / f_sum
            closest_harmonic = harmonic
        end
    end

    return (closest_period, closest_harmonic)
end

f_annual = 1 / 365.25
f_moon = 1 / 13.606
target_period = 433   ##!!!! fix this, make it floating-point 433.0
max_harmonic = 50

closest_period, closest_harmonic = find_closest_aliased_harmonic(f_annual, f_moon, target_period, max_harmonic)
println("Closest aliased harmonic: ", closest_harmonic)
println("Closest period: ", closest_period, " days")

This function find_closest_aliased_harmonic takes the annual and Moon’s frequencies (f_annual and f_moon), the target period (433 days), and the maximum number of harmonics to consider (max_harmonic). It iterates through different harmonics of the annual cyclic impulse and calculates the sum and difference frequencies with the Moon’s cycle. The function returns the closest harmonic and its corresponding period.

In this example, I’ve set the maximum number of harmonics to 50. You can adjust the max_harmonic parameter to include more or fewer harmonics, depending on the desired level of precision.

Keep in mind that this is a simplified analysis, and other factors might influence the Earth’s rotation and wobble. Moreover, this algorithm assumes that the primary aliased harmonic comes from the interaction between the annual cyclic impulse and the Moon’s cycle. Other celestial bodies or internal processes might contribute to the wobble, and their effects are not considered in this analysis.

julia> closest_period, closest_harmonic = find_closest_aliased_harmonic(f_annual, f_moon, target_period, max_harmonic)
(432.36397250739594, 26)
julia> println("Closest aliased harmonic: ", closest_harmonic)
Closest aliased harmonic: 26
julia> println("Closest period: ", closest_period, " days")
Closest period: 432.36397250739594 days






The bottom-line response !!!

Closest period: 432.36397250739594 days

Had to make one fix, replacing 433 with 433.0 in the target_period above. Everything ran fine but had to increase the forcing amplitude to 10 from 0.01 to see the forced response above the natural resonance.

I also asked AgentGPT to execute a plan for solving the Chandler Wobble

Sub(Surface)Stack

I signed up for a SubStack account awhile ago and recently published two articles on this account (SubSurface) in the last week.

The SubStack authoring interface has good math equation mark-up, convenient graphics embedding, and an excellent footnoting system. On first pass, it only lacks control over font color.

The articles are focused on applying neural network cross-validation to ENSO and AMO modeling, as suggested previously. I haven’t completely explored the configuration space but one aspect that may becoming clear is the value of wavelet neural networks (WNN) for time-series analysis. The WNN approach seems much more amenable to extracting sinusoidal modulation of the input-to-output mapping — trained on a rather short interval and then cross-validated out-of-band. The Mexican hat wavelet (2nd derivative of a Gaussian) as an activation function in particular locks in quickly to an LTE modulation that took longer to find with the custom search software I have developed at GitHub. I think the reason for the efficiency is that it’s optimizing to a Taylor’s series expansion of the input terms, a classic nonlinear expansion that NN’s excel at.

The following training run using the Mexican hat activation and ADAM optimizer is an eye-opener, as it achieved an admirable fit within a minute of computation.

The GREEN on BLUE is training on NINO4 data over two end-point intervals, with the RED cross-validation over the out-of-band region. The correlation coefficient is 0.34, which is impressive considering the nature of the waveform. Clearly there is similarity.

Moreover, if we compare the model fit to data via the WNN against the LTE harmonics approach, you can also see where the two fare equally poorly. Below in the outer frame is the NINO4 LTE fit with the YELLOW arrow pointing downward at a discrepancy (a peak in the data not resolved in the fit). In comparison the yellow-bordered inset shows the same discrepancy on the WNN training run. So the fingerprints essentially match with no coaching.

The neural net chain is somewhat deep with 6 layers, but I think this is needed to expand to the higher-order terms in the Taylor’s series. In the directed graph below, L01 is the input tidal forcing and L02 is the time axis (with an initial very low weighting).

It also appears temporally stationary across the entire time-span, so that the WNN temporal contribution appears minimal.

In a previous fit the horizontal striations (indicating modulation factor at a forcing level) matched with the LTE model, providing further evidence that the the WNN was mapping to an optimal modulation.

The other Sub(Surface)Stack article is on the AMO, which also reveals promising results. This is a video of the training in action

Intersectionality and No Regrets

The issue is Intersectionality and the solution is a No Regrets Strategy, which means that you don’t regret the choice you make because at least some of the bases are covered.

After filling in the blanks of the Venn diagram:

Quick YT explainer of No Regrets

Interesting to ask GPT-4 what No Regrets means in the context of the earth, i.e. ContextEarth.com

What does the “No Regrets Strategy” mean in terms of mitigating climate change, fossil fuel depletion, and air pollution?

The “No Regrets Strategy” in the context of climate change mitigation, fossil fuel depletion, and air pollution refers to a set of policies, measures, and actions that provide multiple benefits, regardless of the uncertainties associated with the future development of these issues. These measures are seen as advantageous and worthwhile, even if the worst predictions of climate change, fossil fuel depletion, or air pollution do not materialize.

The No Regrets Strategy typically includes:

  1. Energy efficiency improvements: By improving energy efficiency, we can reduce energy consumption, decrease greenhouse gas emissions, reduce dependence on fossil fuels, and lower air pollution. These improvements have economic benefits, as they lead to cost savings for businesses and consumers.
  2. Renewable energy development: Investing in renewable energy sources, such as solar, wind, and hydro, helps to diversify energy supply, reduce greenhouse gas emissions, and decrease reliance on fossil fuels. This transition also has the potential to create new jobs and stimulate economic growth.
  3. Afforestation and reforestation: Planting trees and restoring forests can help to sequester carbon dioxide, enhance biodiversity, and improve air and water quality. This strategy also helps to protect ecosystems and provides valuable recreational spaces for communities.
  4. Sustainable transportation: Encouraging public transportation, walking, cycling, and the use of electric vehicles reduces fossil fuel consumption, greenhouse gas emissions, and air pollution. These measures can also help to alleviate traffic congestion and improve public health.
  5. Waste reduction and recycling: Reducing waste generation and promoting recycling conserves resources, reduces greenhouse gas emissions, and lowers air and water pollution. These actions can also create jobs and generate economic benefits.

By implementing these no-regrets measures, societies can mitigate climate change, reduce fossil fuel dependence, and improve air quality, while also realizing economic, social, and environmental benefits.


The bible of the No Regrets Strategy is Mathematical GeoEnergy


PostScript:Don’t know if this is the intent but this cartoon sums up the No Regrets strategy perfectly