Dynamics II
Lecture: May 12, 2025 (Monday), 14:00 Prof. Dr. Gerrit Lohmann
Linear stability analysis
Consider the continuous dynamical system described by \[\dot x=f(x,\lambda)\quad \] A bifurcation occurs at \[(x_E,\lambda_0)\] if the Jacobian matrix \[ \textrm{d}f/dx (x_E,\lambda_0)\] has an Eigenvalue with zero real part.
Galerkin approximation: Get a low-order model
\[ \mbox{ Saltzman (1962): Expand } \Psi, \Theta \mbox{ in double Fourier series in x and z: } \]
\[\Psi (x,z,t) \, = \, \sum_{k=1}^\infty \sum_{l=1}^\infty \Psi_{k,l} (t) \, \, \sin \left(\frac{k \pi a}{H} x \right) \, \times \, \sin \left(\frac{ l \pi}{H} z \right) \] \[\Theta (x,z,t) \, = \, \sum_{k=1}^\infty \sum_{l=1}^\infty \Theta_{k,l} (t) \cos \left(\frac{k \pi a}{H} x \right) \, \times \, \sin \left( \frac{l \pi}{H} z \right) \]
From partial differential equations to ordinary differential equations for \(\Psi_{k,l}(t)\) and \(\Theta_{k,l}(t)\).
Approximation: Just 3 Modes X(t), Y(t), Z(t)
\[ \frac{a}{1+a^2} \, \kappa \, \Psi = X \sqrt{2} \sin\left(\frac{\pi a}{H} x \right) \sin\left(\frac{\pi}{H} z \right) \]
\[ \pi \frac{R_a}{R_c} \frac{1}{\Delta T} \, \Theta = Y \sqrt{2} \cos\left(\frac{\pi a}{H} x\right) \sin\left(\frac{\pi}{H} z \right) - Z \sin\left(2 \frac{\pi}{H} z \right) \]
Rayleigh number Ra: Buoyancy & Viscosity
\[ \mbox{Motion develops if } \quad R_a = \frac{g \alpha H^3 \Delta T}{\nu \kappa} \quad \mbox{exceeds a critical } \quad R_c \]
As the Rayleigh number increases, the gravitational force becomes more dominant. The critical Rayleigh number can be obtained analytically for a number of different boundary conditions by doing a perturbation analysis on the linearized equations in the stable state.
\(R_c = \pi^4 \frac{\left(1+a^2\right)^3}{a^2} = \pi^4 \frac{27}{4}= 657.51\) with \(a^2 = 1/2\)
\[\mbox{When } R_a < R_c,\mbox{ heat transfer is due to conduction} \]
\[\mbox{When } R_a > R_c, \mbox{ heat transfer is due to convection.} \]
Lorenz system
Bifurcation at \[ r = R_a/R_c = 1\]
Famous low-order model:
\[ \dot X = -\sigma X + \sigma Y \]
\[ \dot Y = r X - Y - X Z \]
\[ \dot Z = -b Z + X Y \]
North Atlantic Current & Gulfstream
brings warm water northward where it cools.
returns southward as a cold, deep, western-boundary current.
Gulf Stream carries 40 Sv of 18°C water northward.
Of this, 15 Sv return southward in the deep western boundary current at a temperature of 2°C.
How much heat is transported northward ?
Calculation:
\[ \underbrace{ c_p}_{4.2 \cdot 10^3 Ws/(m^3 kg)} \, \cdot \, \underbrace{ \rho }_{10^3 kg/m^3 } \, \cdot \, \underbrace{\Phi}_{15 \cdot 10^6 m^3/s} \, \cdot \, \underbrace{\Delta T}_{(18-2) K } = 1 \cdot 10^{15} W \]
The flow carried by the conveyor belt loses 1 Petawatts (PW), close to estimates of Rintoul and Wunsch (1991)
The deep bottom water from the North Atlantic is mixed upward in other regions and ocean, and it makes its way back to the Gulf Stream and the North Atlantic. Thus most of the water that sinks in the North Atlantic must be replaced by water from the far South Atlantic and Pacific Ocean.
Ocean Conveyor Belt
Conveyor Belt: Industry
Conveyor belt circulation
The the conveyor is driven by deepwater formation in the northern North Atlantic.
The conveyor belt metaphor necessarily simplifies the ocean system, it is of course not a full description of the deep ocean circulation.
Broecker’s concept provides a successful approach for global ocean circulation, although several features can be wrong like the missing Antarctic bottom water, the upwelling areas etc..
metaphor inspired new ideas of halting or reversing the ocean circulation and put it into a global climate context.
interpretation of Greenland ice core records indicating different climate states with different ocean modes of operation (like on and off states of a mechanical maschine).
Thermohaline ocean circulation
Modelled meridional overturning streamfunction in Sv 10^6 = m^3 /s in the Atlantic Ocean. Grey areas represent zonally integrated smoothed bathymetry
Estimates of overturning ?
It is observed that water sinks in to the deep ocean in polar regions of the Atlantic basin at a rate of 15 Sv. (Atlantic basin: 80,000,000 km^2 area * 4 km depth.)
– How long would it take to ‘fill up’ the Atlantic basin?
– Supposing that the local sinking is balanced by large-scale upwelling, estimate the strength of this upwelling.
Hint: Upwelling = area * w
– Compare this number with that of the Ekman pumping!
Estimates of overturning: Solution
Timescale T to ‘fill up’ the Atlantic basin: \[ T = \frac{ 80 \cdot 10^{12} \, m^2 \cdot 4000 \, m}{15 \cdot 10^6 \, m^3 s^{-1}} = 2.13 \cdot 10^{10} s = 676 \;years\]
Overturning is balanced by large-scale upwelling:
\[ area \cdot w = 15 \cdot 10^6 \, m^3 s^{-1}\]
\[ w = 0.1875 \cdot 10^{-6} m\;s^{-1} = 5.9 \cdot 10^{-15} m \; y^{-1}. \]
Ekman pumping \[ w_E \simeq 32 \, \, m \; y^{-1}. \]
Stability of the 0-D Energy Balance Model
The zero-dimensional energy balance model (EBM) with ice-albedo feedback is given by: \[ C \frac{dT}{dt} = \frac{S_0}{4} (1 - \alpha(T)) - \epsilon \sigma T^4 \]
Where:
- \(T\): Global mean surface
temperature
- \(C\): Effective heat capacity of
the climate system
- \(S_0 \approx 1361 \,
\text{W/m}^2\): Solar constant
- \(\alpha(T)\): Planetary albedo,
dependent on temperature
- \(\epsilon \approx 0.61\):
Effective emissivity
- \(\sigma = 5.67 \times 10^{-8} \, \text{W/m}^2\text{K}^4\): Stefan–Boltzmann constant
Ice-Albedo Feedback
The albedo is parameterized as: \[ \alpha(T) = \alpha_{\text{ice}} + \frac{\alpha_{\text{water}} - \alpha_{\text{ice}}}{1 + \exp\left( \frac{T - T_c}{\Delta T} \right)} \] with typical values:
- \(\alpha_{\text{ice}} = 0.6\):
Albedo of ice
- \(\alpha_{\text{water}} = 0.3\):
Albedo of open water
- \(T_c = 273\,\text{K}\): Critical
temperature
- \(\Delta T = 5\,\text{K}\): Transition width
Equilibrium and Stability
At equilibrium: \[ \frac{S_0}{4} (1 - \alpha(T)) = \epsilon \sigma T^4 \]
The function: \[ f(T) = \frac{S_0}{4}(1 - \alpha(T)) - \epsilon \sigma T^4 \]
Stability is determined by the sign of the derivative: \[ f'(T) = -\frac{S_0}{4} \alpha'(T) - 4\epsilon \sigma T^3 \]
If \(f'(T_e) < 0\), the equilibrium at \(T_e\) is stable; if \(f'(T_e) > 0\), it is unstable.
# Load necessary libraries
library(ggplot2)
library(gridExtra)
# Physical constants
S0 <- 1361 # Solar constant [W/m^2]
epsilon <- 0.61 # Emissivity
sigma <- 5.67e-8 # Stefan–Boltzmann constant [W/m^2/K^4]
# Albedo model parameters
alpha_ice <- 0.7
alpha_water <- 0.3
T_c <- 260 # Transition temperature [K]
delta_T <- 5 # Transition width [K]
# Albedo and its derivative
alpha <- function(T) {
alpha_water + (alpha_ice - alpha_water) / (1 + exp((T - T_c) / delta_T))
}
alpha_prime <- function(T) {
expT <- exp((T - T_c) / delta_T)
- (alpha_ice - alpha_water) * expT / ( delta_T * (1 + expT)^2)
}
onemalpha_int <- function(T) {
expT <- exp((T - T_c) / delta_T)
(1-alpha_water) * T - (alpha_ice - alpha_water) * delta_T * log( expT/(1 + expT))
}
# Energy balance and its derivative
f_T <- function(T) {
(S0 / 4) * (1 - alpha(T)) - epsilon * sigma * T^4
}
f_prime <- function(T) {
- (S0 / 4) * alpha_prime(T) - 4 * epsilon * sigma * T^3
}
f_int <- function(T) {
-(S0 / 4) * onemalpha_int(T) + epsilon/5 * sigma * T^5
}
# Temperature range and values
T_vals <- seq(220, 310, by = 0.1)
f_vals <- sapply(T_vals, f_T)
f_prime_vals <- sapply(T_vals, f_prime)
f_int_vals <- sapply(T_vals, f_int)
# Find approximate equilibria (sign changes of f)
signs <- sign(f_vals)
eq_indices <- which(diff(signs) != 0)
T_eq <- T_vals[eq_indices]
f_eq <- sapply(T_eq, f_T)
f_prime_eq <- sapply(T_eq, f_prime)
stability <- ifelse(f_prime_eq < 0, "Stable", "Unstable")
# Create data frames
df_f <- data.frame(T = T_vals, f = f_vals)
df_fp <- data.frame(T = T_vals, f_prime = f_prime_vals)
df_fint <- data.frame(T = T_vals, f_int = f_int_vals)
df_eq <- data.frame(T = T_eq, f = f_eq, f_prime = f_prime_eq, Stability = stability)
# Plot f(T)
plot_f <- ggplot(df_f, aes(x = T, y = f)) +
geom_line(color = "green", size = 1) +
geom_hline(yintercept = 0, linetype = "dashed") +
geom_point(data = df_eq, aes(x = T, y = f, color = Stability), size = 3) +
scale_color_manual(values = c("Stable" = "blue", "Unstable" = "orange")) +
labs(title = "Net Energy Flux f(T)", x = "Temperature [K]", y = expression(f(T)~"[W/m"^2*"]")) +
theme_minimal()
# Plot f'(T)
plot_fp <- ggplot(df_fp, aes(x = T, y = f_prime)) +
geom_line(color = "red", size = 1) +
geom_hline(yintercept = 0, linetype = "dashed") +
geom_point(data = df_eq, aes(x = T, y = f_prime, color = Stability), size = 3) +
scale_color_manual(values = c("Stable" = "blue", "Unstable" = "orange")) +
labs(title = "Derivative f'(T)", x = "Temperature [K]", y = expression(f*"'(T)~[W/m"^2*"/K]")) +
theme_minimal()
# Plot integral of f(T)
plot_fint <- ggplot(df_fint, aes(x = T, y = f_int)) +
geom_line(color = "red", size = 1) +
geom_hline(yintercept = 0, linetype = "dashed") +
geom_point(data = df_eq, aes(x = T, y = f_prime, color = Stability), size = 3) +
scale_color_manual(values = c("Stable" = "blue", "Unstable" = "orange")) +
labs(title = "Integral f(T)", x = "Temperature [K]", y = expression(V(T)~"[W/m"^2*"K]")) +
theme_minimal()
# Combine the two plots
grid.arrange(plot_f, plot_fp, ncol = 2)
Scaling: Rotating frame of reference
The Coriolis effect is one of the dominating forces for the large-scale dynamics.
\[ \underbrace{\frac{\partial \mathbf{v}}{\partial t}}_{ U/T \sim 10^{-8} } + \underbrace{\mathbf{v} \cdot \nabla \mathbf{v}}_{ U^2/L \sim 10^{-8} }= {\underbrace{- \frac{1}{\rho} \nabla p}_{ \bf \delta P/(\rho L) \sim 10^{-5} } + \underbrace{2 \mathbf{\Omega \times v}}_{ \bf f_0 U \sim 10^{-5} } + \underbrace{fric}_{ \nu U/H^2 \sim 10^{-13}}} \quad \]
where fric is due to eddy stress divergence (\(\sim \nu \nabla^2 \mathbf{v}\)).
Values from the ocean -> exercise.
Because of the continuity equation \[ U/L \sim W/H \] and horizontal scales are orders of magnitude larger than the vertical ones, \[ W << U . \]
The timescales are related to \[ T \sim L/U \sim H/W \]
Exception for small scales (e.g., ocean convection or cumuls clouds): \[ H \sim L \quad \rightarrow \quad W \sim U \]
Rossby number Ro
\[ \underbrace{\frac{\partial \mathbf{v}}{\partial t}}_{ U/T \sim 10^{-8} } + \underbrace{\mathbf{v} \cdot \nabla \mathbf{v}}_{ U^2/L \sim 10^{-8} } = {\underbrace{- \frac{1}{\rho} \nabla p}_{ \bf \delta P/(\rho L) \sim 10^{-5}} + \underbrace{2 \mathbf{\Omega \times v}}_{ \bf f_0 U \sim 10^{-5} } + \underbrace{fric}_{ \nu U/H^2 \sim 10^{-13}}} \]
\[ Ro = \frac{ \mbox{Inertial (the left hand side)} }{ \mbox{Coriolis term } } \]
\[ Ro = \frac{(U^2/L)}{(f U)} = \frac{U}{f L} \quad \]
characterizes the importance of Coriolis acceleration
Ro is small when the flow is in a so-called geostrophic balance.
Vorticity is the rotation of the fluid
\[ \zeta \equiv \frac{\partial v}{\partial x}-\frac{\partial u}{\partial y} \]
or in 3D:
\[ \equiv \nabla \times \boldsymbol{ u} \]
Example: Rigid body rotating
\[ \boldsymbol{ u} = \begin{pmatrix} u \\ v \\ w \end{pmatrix}\ = \boldsymbol{ \Omega} \times \boldsymbol{ r} = \begin{pmatrix} \omega_1 \\ \omega_2 \\ \omega_3 \end{pmatrix}\ \times \ \begin{pmatrix} x \\ y \\ z \end{pmatrix}\ = \begin{pmatrix} \omega_2 z- \omega_3 y\\ \omega_3 x- \omega_1 z \\ \omega_1 y - \omega_2 x \end{pmatrix}\ \]
Rotation vector
\[ \nabla \times \boldsymbol{ u} = \begin{pmatrix} \partial_x \\ \partial_y \\ \partial_z \end{pmatrix}\ \times \ \begin{pmatrix} u \\ v \\ w \end{pmatrix}\ = \begin{pmatrix} \partial_x \\ \partial_y \\ \partial_z \end{pmatrix}\ \times \ \begin{pmatrix} \omega_2 z- \omega_3 y \\ \omega_3 x- \omega_1 z \\ \omega_1 y - \omega_2 x \end{pmatrix}\ \]
\[ = \begin{pmatrix} \partial_y (\omega_1 y - \omega_2 x) - \partial_z (\omega_3 x- \omega_1 z ) \\ \\ \\ \end{pmatrix}\ \]
\[ = \begin{pmatrix} \omega_1 + \omega_1 \\ \omega_2 + \omega_2 \\ \omega_3 + \omega_3 \end{pmatrix}\ = 2 \boldsymbol{ \Omega} \]
Example: Vorticity from shear
Tomczak & Godfrey: Regional Oceanography
\[u=0, v=v\left(x\right)\]
\[\zeta= \partial v\left(x\right)/\partial x \]
Estimate for \(\zeta\) off Cape Hatteras:
the velocity changes by \(1 \, {m}{s}^{-1}\) in 100 km
\[ \zeta= \frac{\partial v}{\partial x} = \frac{ 1 \, {m}{s}^{-1}}{100 \, {km}} = 10^{-5} \, \frac{1}{s} \]
still much smaller than
\[ f= 2 \Omega \sin \varphi = 2 \, \frac{2 \pi}{day} \sin \varphi \approx \, 10^{-4} \, \frac{1}{s} \]
Planetary and relative vorticity
\[ \mbox{Absolute Vorticity }\equiv\left(\zeta+f\right) \]
\[ \frac{Du}{Dt}-f\;v = -\frac{1}{\rho}\frac{\partial p}{\partial x} \] \[ \frac{Dv}{Dt}+f\;u = -\frac{1}{\rho}\frac{\partial p}{\partial y} \]
\[ \mbox{subtract } \partial/\partial y \mbox{ of (u-equation) from } \partial /\partial x \mbox{ of (v-equation) } \]
Use \[ \frac{D}{Dt} f = v \, \partial_y f: \]
to obtain
\[ \underline{ \frac{D}{Dt}\left(\zeta+f\right) + \left(\zeta + f\right)\left(\frac{\partial u}{\partial x}+\frac{\partial v}{\partial y}\right)=0 } \quad \]
Examples for Vorticity: Ocean with depth h(x,y)
\[ \mbox{Because of the continuity equation } \quad \partial_x \left( u h \right) + \partial_y \left( v h \right) \quad = \quad 0 \]
\[ \quad \frac{D}{Dt} h + h \left( \partial_x u + \partial_y v \right) = \quad 0 \]
Therefore, \[ \underline{ \frac{D}{Dt}\left(\zeta+f\right) + \left(\zeta + f\right)\left(\frac{\partial u}{\partial x}+\frac{\partial v}{\partial y}\right)=0 } \quad \]
\[ \frac{D}{Dt}\left(\zeta +f\right)-\frac{\left(\zeta+f\right)}{h}\frac{Dh}{Dt}=0 \]
\[ \frac{1}{h} \frac{D}{Dt}\left(\zeta+f\right) - \left(\zeta + f\right) \frac{D_t h}{h^2} =0 \]
\[ \underline{ \frac{D}{Dt}\left( \frac{\zeta+f}{h}\right) = 0 } \quad \]
Potential vorticity is conserved along a fluid trajectory.
Potential vorticity: Examples
\[ \frac{D}{Dt}\left(\zeta+f\right) + \left(\zeta + f\right)\left(\frac{\partial u}{\partial x}+\frac{\partial v}{\partial y}\right)=0 \quad \]
Ocean/Atmosphere with depth h(x,y)
\[ \frac{D}{Dt}\left( \frac{\zeta+f}{h}\right) = 0 \quad \]
Couples depth, vorticity, latitude
– Changes in the depth results in change in \(\zeta\).
– Changes in latitude require a corresponding change in \(\zeta\).
Dietrich et al. (1980)
Steward, Oceanography
Angular Momentum and Hadley Cell
Tropical air rises to tropopause & moves poleward
Deflected eastward by the Coriolis force
Subtropical jet: forms at poleward limit of Hadley Cell
It tends to conserve angular momentum, friction small
equatorward moving air: westward component
What drives the ocean currents?
Friction: transfer of momentum from atmosphere to oceanic Ekman layer
Vorticity dynamics for the ocean and include the wind stress term
\[ D_t u - f v = - \frac{1}{\rho} \frac{\partial p}{\partial x} + \frac{1}{\rho} \partial_z \tau_{xz} \] \[ D_t v + f u = - \frac{1}{\rho} \frac{\partial p}{\partial y} + \frac{1}{\rho} \partial_z \tau_{yz} \]
\[ \frac{D}{Dt} \left( {\zeta+f}\right) - \frac{\left(\zeta+f \right)}{h} \frac{D}{Dt} h \, = \, \frac{1}{\rho} \underbrace{\left( \frac{\partial}{\partial x} \, \partial_z \tau_{yz} - \frac{\partial}{\partial y}\, \partial_z \tau_{xz} \right)}_{curl \, \partial_z \tau} \quad . \]
\[ \frac{D}{Dt} \left( \frac{\zeta+f}{h}\right) = \frac{1}{\rho \, h} \, \mbox{curl} \, \partial_z \tau \, \]
Sverdrup transport
\[ \beta v = \frac{1}{\rho } \, \mbox{curl} \, \partial_z \tau \, \]
\[ \int_{-H}^0 dz \, \beta v = \frac{1}{\rho } \, \int_{-H}^0 dz \, \mbox{curl} \, \partial_z \tau \, = \frac{1}{\rho } \, \mbox{curl} \, \tau \, \]
\[ V = \frac{1}{\rho \beta} \, \left( \frac{\partial \tau_{yz} }{\partial x} \, - \frac{\partial \tau_{xz}}{\partial y}\, \right) = \frac{1}{\rho \beta} \, \, \operatorname{curl} \, \tau \]
applied globally using the wind stress from Hellerman and Rosenstein (1983). Contour interval is \(10\) Sverdrups (Tomczak and Godfrey, 1994).
Ekman Pumping & Sverdrup Transport
The center of a subtropical gyre is a high pressure zone: clockwise on the Northern Hemisphere
Ekman surface currents towards the center of the gyre
The Ekman vertical velocity balanced by \[ w_E=w_g \] vertical geostrophic current in the interior
geostrophic flow towards the equator
returned flow towards the pole in western boundary currents
Ekman Pumping: vertical velocity at the bottom of the Ekman layer E
\(w_E\) as the Ekman vertical velocity the bottom of the Ekman layer \[ w_E = - \int_{-E}^0 \frac{\partial w}{\partial z} dz = \frac{\partial}{\partial x} U_E + \frac{\partial}{\partial y} V_E \]
\(\operatorname{curl} \mathbf{\tau}\) produces a divergence of the Ekman transports leading to \(w_E\) at the bottom E
\[ w_E = \, \frac{\partial }{\partial x} \left( \frac{ \tau_{y}}{\rho \;f }\, \right) - \frac{\partial }{\partial y}\, \left( \frac{ \tau_{x}}{\rho \;f }\, \right) =\operatorname{curl}\left(\frac{\mathbf{\tau}}{\rho\;f}\right) \simeq \frac{1}{\rho\;f} \, \operatorname{curl} \mathbf{\tau} \]
The order of magnitude of the Ekman vertical velocity:
typical wind stress variation of \(0.2 N m^{-2}\) per 2000 km in y-direction:
\[ w_E \simeq - \frac{ \Delta \tau_{x}}{\rho \;f_0 \Delta y}\, \simeq \frac{1 }{10^3 kg m^{-3}} \frac{0.2 N m^{-2} }{10^{-4} s^{-1}\, \, 2 \cdot 10^6 m} \simeq 32 \, \, \frac{m}{yr} \]
Conveyor belt: climate change
left: 55%
halting or reversing the ocean circulation
interpretation of Greenland ice core records
climate states with different ocean modes
Overturning Circulation (model)
Modelled meridional overturning streamfunction in Sv 10^6 =
m^3 /s in the Atlantic Ocean. Grey areas represent zonally integrated
smoothed bathymetry
Atlantic water masses
North
Estimates of overturning
Timescale T to ‘fill up’ the Atlantic basin:
\[ T = \frac{ 80 \cdot 10^{12} \, m^2 \cdot 4000 \, m}{15 \cdot 10^6 \, m^3 s^{-1}} = 2.13 \cdot 10^{10} s = 676 \;years\]
Overturning is balanced by large-scale upwelling:
\[ area \cdot w = 15 \cdot 10^6 \, m^3 s^{-1}\]
\[ w = 0.1875 \cdot 10^{-6} m\;s^{-1} = 5.9 \cdot 10^{-15} m \; y^{-1}. \]
compare to Ekman pumping \[ w_E \simeq 32 \, \, m \; y^{-1}. \]
Vorticity dynamics of meridional overturning (y,z)
\[ \frac{\partial}{\partial t} v \quad = \quad - \frac{1}{\rho_0} \frac{\partial p}{\partial y} \quad - \quad f u \quad - \quad \kappa v \]
\[ \frac{\partial}{\partial t} w \quad = \quad - \frac{1}{\rho_0} \frac{\partial p}{\partial z} \quad - \quad \frac{g}{\rho_0} (\rho -\rho_0) \quad - \quad \kappa w \] \(\kappa\) as parameter for Rayleigh friction.
Amplitude of overturning
\[ \frac{d}{dt} \Phi_{max} \, = \, \frac{a}{\rho_0} (\rho_{north} - \rho_{south}) \, \, - \, \, \kappa \Phi_{max} \] with \[ a = g L H^2/4(L^2 + H^2) \, \]
This shows that the overturning circulation depends on the density differences on the right and left boxes.
It is simplified to a diagnostic relation
\[ \Phi_{max} = \frac{a}{\rho_0 \, \kappa} \, \, (\rho_{north} - \rho_{south}) \quad \]
because the adjustment of \(\Phi_{max}\) is quasi-instantaneous due to adjustment processes, e.g. Kelvin waves.
North-south density gradient in an ocean basin
Primary north-south gradient in balance with an eastward geostrophic current: generates a secondary high & low pressure system, northward current
=======================================================
Schematic picture of the hemispheric two box model (a) and of the interhemispheric box model
=======================================================
- The Atlantic surface density is mainly related to temperature
differences.
- But the pole-to-pole differences are caused by salinity differences. }
Box model of the thermohaline circulation
hemispheric (Stommel-type) or interhemispheric (Rooth-type)
Box model of the thermohaline circulation
hemispheric (Stommel-type) or interhemispheric (Rooth-type)
\[ \Phi \, = \, - c \, ( \alpha \Delta T - \beta \Delta S ) \qquad , \] where \(\alpha\) and \(\beta\) are the thermal and haline expansion coefficients,
\(\Delta\) denotes the meridional difference operator applied to temperature T and salinity S, respectively.
single hemispheric view:
The meridional density differences are clearly dominated by temperature
Salinity difference breakes the temperature difference
both hemispheres:
densities at high northern and southern latitudes are close,
the pole-to-pole differences are caused by salinity differences.
Ocean: multiple equilibria in GCM
Ocean: multiple equilibria in GCM
Ocean response: Hosing the North Atlantic
- AMOC indices of North Atlantic hosing for different hosing areas. Units are Sv. Black line represents the unperturbed LGM experiment. Hosing is for the period 840–990. (b) Annual mean sea surface salinity anomaly between LGM and the perturbation experiment LGM with 0.2 Sv for the model years 900–950.
-> Multi-scale Ocean GCM
Conveyor belt: climate change
left: 55%
halting or reversing the ocean circulation
interpretation of Greenland ice core records
climate states with different ocean modes
Abrupt climate change, termination of ice sheets, Climate System II
Ocean box model
Ocean: Temperature
Euler forward for ocean temperature
Tnln = Tnl + dts * ((Hfnl)/(rcz2)-(Tnl-Tml)*phi/Vnl);
Tmln = Tml + dts * ((Hfml)/(rcz1)-(Tml-Tsl)*phi/Vml);
Tsln = Tsl + dts * ((Hfsl)/(rcz2)-(Tsl-Td)*phi/Vsl);
Tdn = Td + dts * (-(Td-Tnl)*(phi/Vd));
Atmosphere
Box model
3) Boxmodel (Version using R-shiny)
How to setup and run the Boxmodel R-shiny app
password: DynamicsII2020
Box model with jupyter
for the box model, you can download sevenbox_jupyter.ipynb
conda create -n jupyter-R
conda activate jupyter-R
conda install -y -c conda-forge pip notebook nb_conda_kernels jupyter_contrib_nbextensions
conda install -y -c conda-forge r r-irkernel r-ggplot2 r-dplyr
jupyter-notebook sevenbox_jupyter.ipynb
you have to download conda (or miniconda). Here a good web site: https://conda.io/projects/conda/en/latest/user-guide/install/index.html
Boxmodel with R-shiny
Download link password: DynamicsII2020
go to the folder where the boxmodel is located (ui.R): setwd(’ … ’)
install.packages(“shiny”)
library(shiny)
runApp()
a browser window should open that displays the boxmodel controls
first click on the button Spin-Up to create a new spin up
a new window opens, click on Run SpinUp
afterwards click on Apply, then Cancel
define parameters for the model simulation
run the simulation by clicking on Start Simulation