Introductory theory

This tutorial was generated using Literate.jl. Download the source as a .jl file. Download the source as a .ipynb file.

Note

This tutorial is aimed at advanced undergraduates or early-stage graduate students. You don't need prior exposure to stochastic programming! (Indeed, it may be better if you don't, because our approach is non-standard in the literature.)

This tutorial is also a living document. If parts are unclear, please open an issue so it can be improved!

This tutorial will teach you how the stochastic dual dynamic programming algorithm works by implementing a simplified version of the algorithm.

Our implementation is very much a "vanilla" version of SDDP; it doesn't have (m)any fancy computational tricks (e.g., the ones included in SDDP.jl) that you need to code a performant or stable version that will work on realistic instances. However, our simplified implementation will work on arbitrary policy graphs, including those with cycles such as infinite horizon problems!

Packages

This tutorial uses the following packages. For clarity, we call import PackageName so that we must prefix PackageName. to all functions and structs provided by that package. Everything not prefixed is either part of base Julia, or we wrote it.

import ForwardDiff
import HiGHS
import JuMP
import Statistics
Tip

You can follow along by installing the above packages, and copy-pasting the code we will write into a Julia REPL. Alternatively, you can download the Julia .jl file which created this tutorial from GitHub.

Preliminaries: background theory

Start this tutorial by reading An introduction to SDDP.jl, which introduces the necessary notation and vocabulary that we need for this tutorial.

Preliminaries: Kelley's cutting plane algorithm

Kelley's cutting plane algorithm is an iterative method for minimizing convex functions. Given a convex function $f(x)$, Kelley's constructs an under-approximation of the function at the minimum by a set of first-order Taylor series approximations (called cuts) constructed at a set of points $k = 1,\ldots,K$:

\[\begin{aligned} f^K = \min\limits_{\theta \in \mathbb{R}, x \in \mathbb{R}^N} \;\; & \theta\\ & \theta \ge f(x_k) + \frac{d}{dx}f(x_k)^\top (x - x_k),\quad k=1,\ldots,K\\ & \theta \ge M, \end{aligned}\]

where $M$ is a sufficiently large negative number that is a lower bound for $f$ over the domain of $x$.

Kelley's cutting plane algorithm is a structured way of choosing points $x_k$ to visit, so that as more cuts are added:

\[\lim_{K \rightarrow \infty} f^K = \min\limits_{x \in \mathbb{R}^N} f(x)\]

However, before we introduce the algorithm, we need to introduce some bounds.

Bounds

By convexity, $f^K \le f(x)$ for all $x$. Thus, if $x^*$ is a minimizer of $f$, then at any point in time we can construct a lower bound for $f(x^*)$ by solving $f^K$.

Moreover, we can use the primal solutions $x_k^*$ returned by solving $f^k$ to evaluate $f(x_k^*)$ to generate an upper bound.

Therefore, $f^K \le f(x^*) \le \min\limits_{k=1,\ldots,K} f(x_k^*)$.

When the lower bound is sufficiently close to the upper bound, we can terminate the algorithm and declare that we have found an solution that is close to optimal.

Implementation

Here is pseudo-code fo the Kelley algorithm:

  1. Take as input a convex function $f(x)$ and a iteration limit $K_{max}$. Set $K = 0$, and initialize $f^K$. Set $lb = -\infty$ and $ub = \infty$.
  2. Solve $f^K$ to obtain a candidate solution $x_{K+1}$.
  3. Update $lb = f^K$ and $ub = \min\{ub, f(x_{K+1})\}$.
  4. Add a cut $\theta \ge f(x_{K+1}) + \frac{d}{dx}f\left(x_{K+1}\right)^\top (x - x_{K+1})$ to form $f^{K+1}$.
  5. Increment $K$.
  6. If $K = K_{max}$ or $|ub - lb| < \epsilon$, STOP, otherwise, go to step 2.

And here's a complete implementation:

function kelleys_cutting_plane(
    # The function to be minimized.
    f::Function,
    # The gradient of `f`. By default, we use automatic differentiation to
    # compute the gradient of f so the user doesn't have to!
    dfdx::Function = x -> ForwardDiff.gradient(f, x);
    # The number of arguments to `f`.
    input_dimension::Int,
    # A lower bound for the function `f` over its domain.
    lower_bound::Float64,
    # The number of iterations to run Kelley's algorithm for before stopping.
    iteration_limit::Int,
    # The absolute tolerance ϵ to use for convergence.
    tolerance::Float64 = 1e-6,
)
    # Step (1):
    K = 0
    model = JuMP.Model(HiGHS.Optimizer)
    JuMP.set_silent(model)
    JuMP.@variable(model, θ >= lower_bound)
    JuMP.@variable(model, x[1:input_dimension])
    JuMP.@objective(model, Min, θ)
    x_k = fill(NaN, input_dimension)
    lower_bound, upper_bound = -Inf, Inf
    while true
        # Step (2):
        JuMP.optimize!(model)
        x_k .= JuMP.value.(x)
        # Step (3):
        lower_bound = JuMP.objective_value(model)
        upper_bound = min(upper_bound, f(x_k))
        println("K = $K : $(lower_bound) <= f(x*) <= $(upper_bound)")
        # Step (4):
        JuMP.@constraint(model, θ >= f(x_k) + dfdx(x_k)' * (x .- x_k))
        # Step (5):
        K = K + 1
        # Step (6):
        if K == iteration_limit
            println("-- Termination status: iteration limit --")
            break
        elseif abs(upper_bound - lower_bound) < tolerance
            println("-- Termination status: converged --")
            break
        end
    end
    println("Found solution: x_K = ", x_k)
    return
end
kelleys_cutting_plane (generic function with 2 methods)

Let's run our algorithm to see what happens:

kelleys_cutting_plane(
    input_dimension = 2,
    lower_bound = 0.0,
    iteration_limit = 20,
) do x
    return (x[1] - 1)^2 + (x[2] + 2)^2 + 1.0
end
K = 0 : 0.0 <= f(x*) <= 6.0
K = 1 : 0.0 <= f(x*) <= 2.25
K = 2 : 0.0 <= f(x*) <= 2.25
K = 3 : 0.0 <= f(x*) <= 1.0986328124999998
K = 4 : 0.0 <= f(x*) <= 1.0986328124999998
K = 5 : 0.0 <= f(x*) <= 1.0986328124999998
K = 6 : 0.31451789700255117 <= f(x*) <= 1.0986328124999998
K = 7 : 0.5251698041461698 <= f(x*) <= 1.0986328124999998
K = 8 : 0.730061698203828 <= f(x*) <= 1.0986328124999998
K = 9 : 0.8034849495130735 <= f(x*) <= 1.028310187109376
K = 10 : 0.9287004555830413 <= f(x*) <= 1.028310187109376
K = 11 : 0.9487595478289129 <= f(x*) <= 1.005714217279734
K = 12 : 0.9823564684073585 <= f(x*) <= 1.005714217279734
K = 13 : 0.9878994829249161 <= f(x*) <= 1.0023277313143473
K = 14 : 0.9948297447399875 <= f(x*) <= 1.0023277313143473
K = 15 : 0.9966035252317597 <= f(x*) <= 1.0002586136309097
K = 16 : 0.9988344662887892 <= f(x*) <= 1.0002586136309097
K = 17 : 0.9994007197804659 <= f(x*) <= 1.0002586136309097
K = 18 : 0.9995654191110326 <= f(x*) <= 1.0000883284156266
K = 19 : 0.9998024915692466 <= f(x*) <= 1.0000370583103833
-- Termination status: iteration limit --
Found solution: x_K = [0.993919804093335, -1.9997007875004429]
Warning

It's hard to choose a valid lower bound! If you choose one too loose, the algorithm can take a long time to converge. However, if you choose one so tight that $M > f(x^*)$, then you can obtain a suboptimal solution. For a deeper discussion of the implications for SDDP.jl, see Choosing an initial bound.

Preliminaries: approximating the cost-to-go term

In the background theory section, we discussed how you could formulate an optimal policy to a multistage stochastic program using the dynamic programming recursion:

\[\begin{aligned} V_i(x, \omega) = \min\limits_{\bar{x}, x^\prime, u} \;\; & C_i(\bar{x}, u, \omega) + \mathbb{E}_{j \in i^+, \varphi \in \Omega_j}[V_j(x^\prime, \varphi)]\\ & x^\prime = T_i(\bar{x}, u, \omega) \\ & u \in U_i(\bar{x}, \omega) \\ & \bar{x} = x, \end{aligned}\]

where our decision rule, $\pi_i(x, \omega)$, solves this optimization problem and returns a $u^*$ corresponding to an optimal solution. Moreover, we alluded to the fact that the cost-to-go term (the nasty recursive expectation) makes this problem intractable to solve.

However, if, excluding the cost-to-go term (i.e., the SP formulation), $V_i(x, \omega)$ can be formulated as a linear program (this also works for convex programs, but the math is more involved), then we can make some progress by noticing that $x$ only appears as a right-hand side term of the fishing constraint $\bar{x} = x$. Therefore, $V_i(x, \cdot)$ is convex with respect to $x$ for fixed $\omega$. (If you have not seen this result before, try to prove it.)

The fishing constraint $\bar{x} = x$ has an associated dual variable. The economic interpretation of this dual variable is that it represents the change in the objective function if the right-hand side $x$ is increased on the scale of one unit. In other words, and with a slight abuse of notation, it is the value $\frac{d}{dx} V_i(x, \omega)$. (Because $V_i$ is not differentiable, it is a subgradient instead of a derivative.)

If we implement the constraint $\bar{x} = x$ by setting the lower- and upper bounds of $\bar{x}$ to $x$, then the reduced cost of the decision variable $\bar{x}$ is the subgradient, and we do not need to explicitly add the fishing constraint as a row to the constraint matrix.

Tip

The subproblem can have binary and integer variables, but you'll need to use Lagrangian duality to compute a subgradient!

Stochastic dual dynamic programming converts this problem into a tractable form by applying Kelley's cutting plane algorithm to the $V_j$ functions in the cost-to-go term:

\[\begin{aligned} V_i^K(x, \omega) = \min\limits_{\bar{x}, x^\prime, u} \;\; & C_i(\bar{x}, u, \omega) + \theta\\ & x^\prime = T_i(\bar{x}, u, \omega) \\ & u \in U_i(\bar{x}, \omega) \\ & \bar{x} = x \\ & \theta \ge \mathbb{E}_{j \in i^+, \varphi \in \Omega_j}\left[V_j^k(x^\prime_k, \varphi) + \frac{d}{dx^\prime}V_j^k(x^\prime_k, \varphi)^\top (x^\prime - x^\prime_k)\right],\quad k=1,\ldots,K \\ & \theta \ge M. \end{aligned}\]

All we need now is a way of generating these cutting planes in an iterative manner. Before we get to that though, let's start writing some code.

Implementation: modeling

Let's make a start by defining the problem structure. Like SDDP.jl, we need a few things:

  1. A description of the structure of the policy graph: how many nodes there are, and the arcs linking the nodes together with their corresponding probabilities.
  2. A JuMP model for each node in the policy graph.
  3. A way to identify the incoming and outgoing state variables of each node.
  4. A description of the random variable, as well as a function that we can call that will modify the JuMP model to reflect the realization of the random variable.
  5. A decision variable to act as the approximated cost-to-go term.
Warning

In the interests of brevity, there is minimal error checking. Think about all the different ways you could break the code!

Structs

The first struct we are going to use is a State struct that will wrap an incoming and outgoing state variable:

struct State
    in::JuMP.VariableRef
    out::JuMP.VariableRef
end

Next, we need a struct to wrap all of the uncertainty within a node:

struct Uncertainty
    parameterize::Function
    Ω::Vector{Any}
    P::Vector{Float64}
end

parameterize is a function which takes a realization of the random variable $\omega\in\Omega$ and updates the subproblem accordingly. The finite discrete random variable is defined by the vectors Ω and P, so that the random variable takes the value Ω[i] with probability P[i]. As such, P should sum to 1. (We don't check this here, but we should; we do in SDDP.jl.)

Now we have two building blocks, we can declare the structure of each node:

struct Node
    subproblem::JuMP.Model
    states::Dict{Symbol,State}
    uncertainty::Uncertainty
    cost_to_go::JuMP.VariableRef
end
  • subproblem is going to be the JuMP model that we build at each node.
  • states is a dictionary that maps a symbolic name of a state variable to a State object wrapping the incoming and outgoing state variables in subproblem.
  • uncertainty is an Uncertainty object described above.
  • cost_to_go is a JuMP variable that approximates the cost-to-go term.

Finally, we define a simplified policy graph as follows:

struct PolicyGraph
    nodes::Vector{Node}
    arcs::Vector{Dict{Int,Float64}}
end

There is a vector of nodes, as well as a data structure for the arcs. arcs is a vector of dictionaries, where arcs[i][j] gives the probability of transitioning from node i to node j, if an arc exists.

To simplify things, we will assume that the root node transitions to node 1 with probability 1, and there are no other incoming arcs to node 1. Notably, we can still define cyclic graphs though!

We also define a nice show method so that we don't accidentally print a large amount of information to the screen when creating a model:

function Base.show(io::IO, model::PolicyGraph)
    println(io, "A policy graph with $(length(model.nodes)) nodes")
    println(io, "Arcs:")
    for (from, arcs) in enumerate(model.arcs)
        for (to, probability) in arcs
            println(io, "  $(from) => $(to) w.p. $(probability)")
        end
    end
    return
end

Functions

Now we have some basic types, let's implement some functions so that the user can create a model.

First, we need an example of a function that the user will provide. Like SDDP.jl, this takes an empty subproblem, and a node index, in this case t::Int. You could change this function to change the model, or define a new one later in the code.

We're going to copy the example from An introduction to SDDP.jl, with some minor adjustments for the fact we don't have many of the bells and whistles of SDDP.jl. You can probably see how some of the SDDP.jl functionality like @stageobjective and SDDP.parameterize help smooth some of the usability issues like needing to construct both the incoming and outgoing state variables, or needing to explicitly declare return states, uncertainty.

function subproblem_builder(subproblem::JuMP.Model, t::Int)
    # Define the state variables. Note how we fix the incoming state to the
    # initial state variable regardless of `t`! This isn't strictly necessary;
    # it only matters that we do it for the first node.
    JuMP.@variable(subproblem, volume_in == 200)
    JuMP.@variable(subproblem, 0 <= volume_out <= 200)
    states = Dict(:volume => State(volume_in, volume_out))
    # Define the control variables.
    JuMP.@variables(subproblem, begin
        thermal_generation >= 0
        hydro_generation >= 0
        hydro_spill >= 0
        inflow
    end)
    # Define the constraints
    JuMP.@constraints(
        subproblem,
        begin
            volume_out == volume_in + inflow - hydro_generation - hydro_spill
            demand_constraint, thermal_generation + hydro_generation == 150.0
        end
    )
    # Define the objective for each stage `t`. Note that we can use `t` as an
    # index for t = 1, 2, 3.
    fuel_cost = [50.0, 100.0, 150.0]
    JuMP.@objective(subproblem, Min, fuel_cost[t] * thermal_generation)
    # Finally, we define the uncertainty object. Because this is a simplified
    # implementation of SDDP, we shall politely ask the user to only modify the
    # constraints, and not the objective function! (Not that it changes the
    # algorithm, we just have to add more information to keep track of things.)
    uncertainty = Uncertainty([0.0, 50.0, 100.0], [1 / 3, 1 / 3, 1 / 3]) do ω
        return JuMP.fix(inflow, ω)
    end
    return states, uncertainty
end
subproblem_builder (generic function with 1 method)

The next function we need to define is the analog of SDDP.PolicyGraph. It should be pretty readable.

function PolicyGraph(
    subproblem_builder::Function;
    graph::Vector{Dict{Int,Float64}},
    lower_bound::Float64,
    optimizer,
)
    nodes = Node[]
    for t in 1:length(graph)
        # Create a model.
        model = JuMP.Model(optimizer)
        JuMP.set_silent(model)
        # Use the provided function to build out each subproblem. The user's
        # function returns a dictionary mapping `Symbol`s to `State` objects,
        # and an `Uncertainty` object.
        states, uncertainty = subproblem_builder(model, t)
        # Now add the cost-to-go terms:
        JuMP.@variable(model, cost_to_go >= lower_bound)
        obj = JuMP.objective_function(model)
        JuMP.@objective(model, Min, obj + cost_to_go)
        # If there are no outgoing arcs, the cost-to-go is 0.0.
        if length(graph[t]) == 0
            JuMP.fix(cost_to_go, 0.0; force = true)
        end
        push!(nodes, Node(model, states, uncertainty, cost_to_go))
    end
    return PolicyGraph(nodes, graph)
end
Main.PolicyGraph

Then, we can create a model using the subproblem_builder function we defined earlier:

model = PolicyGraph(
    subproblem_builder;
    graph = [Dict(2 => 1.0), Dict(3 => 1.0), Dict{Int,Float64}()],
    lower_bound = 0.0,
    optimizer = HiGHS.Optimizer,
)
A policy graph with 3 nodes
Arcs:
  1 => 2 w.p. 1.0
  2 => 3 w.p. 1.0

Implementation: helpful samplers

Before we get properly coding the solution algorithm, it's also going to be useful to have a function that samples a realization of the random variable defined by Ω and P.

function sample_uncertainty(uncertainty::Uncertainty)
    r = rand()
    for (p, ω) in zip(uncertainty.P, uncertainty.Ω)
        r -= p
        if r < 0.0
            return ω
        end
    end
    return error("We should never get here because P should sum to 1.0.")
end
sample_uncertainty (generic function with 1 method)
Note

rand() samples a uniform random variable in [0, 1).

For example:

for i in 1:3
    println("ω = ", sample_uncertainty(model.nodes[1].uncertainty))
end
ω = 50.0
ω = 50.0
ω = 0.0

It's also going to be useful to define a function that generates a random walk through the nodes of the graph:

function sample_next_node(model::PolicyGraph, current::Int)
    if length(model.arcs[current]) == 0
        # No outgoing arcs!
        return nothing
    else
        r = rand()
        for (to, probability) in model.arcs[current]
            r -= probability
            if r < 0.0
                return to
            end
        end
        # We looped through the outgoing arcs and still have probability left
        # over! This means we've hit an implicit "zero" node.
        return nothing
    end
end
sample_next_node (generic function with 1 method)

For example:

for i in 1:3
    # We use `repr` to print the next node, because `sample_next_node` can
    # return `nothing`.
    println("Next node from $(i) = ", repr(sample_next_node(model, i)))
end
Next node from 1 = 2
Next node from 2 = 3
Next node from 3 = nothing

This is a little boring, because our graph is simple. However, more complicated graphs will generate more interesting trajectories!

Implementation: the forward pass

Recall that, after approximating the cost-to-go term, we need a way of generating the cuts. As the first step, we need a way of generating candidate solutions $x_k^\prime$. However, unlike the Kelley's example, our functions $V_j^k(x^\prime, \varphi)$ need two inputs: an outgoing state variable and a realization of the random variable.

One way of getting these inputs is just to pick a random (feasible) value. However, in doing so, we might pick outgoing state variables that we will never see in practice, or we might infrequently pick outgoing state variables that we will often see in practice. Therefore, a better way of generating the inputs is to use a simulation of the policy, which we call the forward pass.

The forward pass walks the policy graph from start to end, transitioning randomly along the arcs. At each node, it observes a realization of the random variable and solves the approximated subproblem to generate a candidate outgoing state variable $x_k^\prime$. The outgoing state variable is passed as the incoming state variable to the next node in the trajectory.

function forward_pass(model::PolicyGraph, io::IO = stdout)
    println(io, "| Forward Pass")
    # First, get the value of the state at the root node (e.g., x_R).
    incoming_state =
        Dict(k => JuMP.fix_value(v.in) for (k, v) in model.nodes[1].states)
    # `simulation_cost` is an accumlator that is going to sum the stage-costs
    # incurred over the forward pass.
    simulation_cost = 0.0
    # We also need to record the nodes visited and resultant outgoing state
    # variables so we can pass them to the backward pass.
    trajectory = Tuple{Int,Dict{Symbol,Float64}}[]
    # Now's the meat of the forward pass: beginning at the first node:
    t = 1
    while t !== nothing
        node = model.nodes[t]
        println(io, "| | Visiting node $(t)")
        # Sample the uncertainty:
        ω = sample_uncertainty(node.uncertainty)
        println(io, "| | | ω = ", ω)
        # Parameterizing the subproblem using the user-provided function:
        node.uncertainty.parameterize(ω)
        println(io, "| | | x = ", incoming_state)
        # Update the incoming state variable:
        for (k, v) in incoming_state
            JuMP.fix(node.states[k].in, v; force = true)
        end
        # Now solve the subproblem and check we found an optimal solution:
        JuMP.optimize!(node.subproblem)
        if JuMP.termination_status(node.subproblem) != JuMP.MOI.OPTIMAL
            error("Something went terribly wrong!")
        end
        # Compute the outgoing state variables:
        outgoing_state = Dict(k => JuMP.value(v.out) for (k, v) in node.states)
        println(io, "| | | x′ = ", outgoing_state)
        # We also need to compute the stage cost to add to our
        # `simulation_cost` accumulator:
        stage_cost =
            JuMP.objective_value(node.subproblem) - JuMP.value(node.cost_to_go)
        simulation_cost += stage_cost
        println(io, "| | | C(x, u, ω) = ", stage_cost)
        # As a penultimate step, set the outgoing state of stage t and the
        # incoming state of stage t + 1, and add the node to the trajectory.
        incoming_state = outgoing_state
        push!(trajectory, (t, outgoing_state))
        # Finally, sample a new node to step to. If `t === nothing`, the
        # `while` loop will break.
        t = sample_next_node(model, t)
    end
    return trajectory, simulation_cost
end
forward_pass (generic function with 2 methods)

Let's take a look at one forward pass:

trajectory, simulation_cost = forward_pass(model);
| Forward Pass
| | Visiting node 1
| | | ω = 100.0
| | | x = Dict(:volume => 200.0)
| | | x′ = Dict(:volume => 0.0)
| | | C(x, u, ω) = 0.0
| | Visiting node 2
| | | ω = 100.0
| | | x = Dict(:volume => 0.0)
| | | x′ = Dict(:volume => 0.0)
| | | C(x, u, ω) = 5000.0
| | Visiting node 3
| | | ω = 50.0
| | | x = Dict(:volume => 0.0)
| | | x′ = Dict(:volume => 0.0)
| | | C(x, u, ω) = 15000.0

Implementation: the backward pass

From the forward pass, we obtained a vector of nodes visited and their corresponding outgoing state variables. Now we need to refine the approximation for each node at the candidate solution for the outgoing state variable. That is, we need to add a new cut:

\[\theta \ge \mathbb{E}_{j \in i^+, \varphi \in \Omega_j}\left[V_j^k(x^\prime_k, \varphi) + \frac{d}{dx^\prime}V_j^k(x^\prime_k, \varphi)^\top (x^\prime - x^\prime_k)\right]\]

or alternatively:

\[\theta \ge \sum\limits_{j \in i^+} \sum\limits_{\varphi \in \Omega_j} p_{ij} p_{\varphi}\left[V_j^k(x^\prime_k, \varphi) + \frac{d}{dx^\prime}V_j^k(x^\prime_k, \varphi)^\top (x^\prime - x^\prime_k)\right]\]

It doesn't matter what order we visit the nodes to generate these cuts for. For example, we could compute them all in parallel, using the current approximations of $V^K_i$.

However, we can be smarter than that.

If we traverse the list of nodes visited in the forward pass in reverse, then we come to refine the $i^{th}$ node in the trajectory, we will already have improved the approximation of the $(i+1)^{th}$ node in the trajectory as well! Therefore, our refinement of the $i^{th}$ node will be better than if we improved node $i$ first, and then refined node $(i+1)$.

Because we walk the nodes in reverse, we call this the backward pass.

Info

If you're into deep learning, you could view this as the equivalent of back-propagation: the forward pass pushes primal information through the graph (outgoing state variables), and the backward pass pulls dual information (cuts) back through the graph to improve our decisions on the next forward pass.

function backward_pass(
    model::PolicyGraph,
    trajectory::Vector{Tuple{Int,Dict{Symbol,Float64}}},
    io::IO = stdout,
)
    println(io, "| Backward pass")
    # For the backward pass, we walk back up the nodes.
    for i in reverse(1:length(trajectory))
        index, outgoing_states = trajectory[i]
        node = model.nodes[index]
        println(io, "| | Visiting node $(index)")
        if length(model.arcs[index]) == 0
            # If there are no children, the cost-to-go is 0.
            println(io, "| | | Skipping node because the cost-to-go is 0")
            continue
        end
        # Create an empty affine expression that we will use to build up the
        # right-hand side of the cut expression.
        cut_expression = JuMP.AffExpr(0.0)
        # For each node j ∈ i⁺
        for (j, P_ij) in model.arcs[index]
            next_node = model.nodes[j]
            # Set the incoming state variables of node j to the outgoing state
            # variables of node i
            for (k, v) in outgoing_states
                JuMP.fix(next_node.states[k].in, v; force = true)
            end
            # Then for each realization of φ ∈ Ωⱼ
            for (pφ, φ) in zip(next_node.uncertainty.P, next_node.uncertainty.Ω)
                # Setup and solve for the realization of φ
                println(io, "| | | Solving φ = ", φ)
                next_node.uncertainty.parameterize(φ)
                JuMP.optimize!(next_node.subproblem)
                # Then prepare the cut `P_ij * pφ * [V + dVdxᵀ(x - x_k)]``
                V = JuMP.objective_value(next_node.subproblem)
                println(io, "| | | | V = ", V)
                dVdx = Dict(
                    k => JuMP.reduced_cost(v.in) for (k, v) in next_node.states
                )
                println(io, "| | | | dVdx′ = ", dVdx)
                cut_expression += JuMP.@expression(
                    node.subproblem,
                    P_ij *
                    pφ *
                    (
                        V + sum(
                            dVdx[k] * (x.out - outgoing_states[k]) for
                            (k, x) in node.states
                        )
                    ),
                )
            end
        end
        # And then refine the cost-to-go variable by adding the cut:
        c = JuMP.@constraint(node.subproblem, node.cost_to_go >= cut_expression)
        println(io, "| | | Adding cut : ", c)
    end
    return nothing
end
backward_pass (generic function with 2 methods)

Implementation: bounds

Lower bounds

Recall from Kelley's that we can obtain a lower bound for $f(x^*)$ be evaluating $f^K$. The analogous lower bound for a multistage stochastic program is:

\[\mathbb{E}_{i \in R^+, \omega \in \Omega_i}[V_i^K(x_R, \omega)] \le \min_{\pi} \mathbb{E}_{i \in R^+, \omega \in \Omega_i}[V_i^\pi(x_R, \omega)]\]

Here's how we compute the lower bound:

function lower_bound(model::PolicyGraph)
    node = model.nodes[1]
    bound = 0.0
    for (p, ω) in zip(node.uncertainty.P, node.uncertainty.Ω)
        node.uncertainty.parameterize(ω)
        JuMP.optimize!(node.subproblem)
        bound += p * JuMP.objective_value(node.subproblem)
    end
    return bound
end
lower_bound (generic function with 1 method)
Note

The implementation is simplified because we assumed that there is only one arc from the root node, and that it pointed to the first node in the vector.

Because we haven't trained a policy yet, the lower bound is going to be very bad:

lower_bound(model)
0.0

Upper bounds

With Kelley's algorithm, we could easily construct an upper bound by evaluating $f(x_K)$. However, it is almost always intractable to evaluate an upper bound for multistage stochastic programs due to the large number of nodes and the nested expectations. Instead, we can perform a Monte Carlo simulation of the policy to build a statistical estimate for the value of $\mathbb{E}_{i \in R^+, \omega \in \Omega_i}[V_i^\pi(x_R, \omega)]$, where $\pi$ is the policy defined by the current approximations $V^K_i$.

function upper_bound(model::PolicyGraph; replications::Int)
    # Pipe the output to `devnull` so we don't print too much!
    simulations = [forward_pass(model, devnull) for i in 1:replications]
    z = [s[2] for s in simulations]
    μ = Statistics.mean(z)
    tσ = 1.96 * Statistics.std(z) / sqrt(replications)
    return μ, tσ
end
upper_bound (generic function with 1 method)
Note

The width of the confidence interval is incorrect if there are cycles in the graph, because the distribution of simulation costs z is not symmetric. The mean is correct, however.

Termination criteria

In Kelley's algorithm, the upper bound was deterministic. Therefore, we could terminate the algorithm when the lower bound was sufficiently close to the upper bound. However, our upper bound for SDDP is not deterministic; it is a confidence interval!

Some people suggest terminating SDDP when the lower bound is contained within the confidence interval. However, this is a poor choice because it is too easy to generate a false positive. For example, if we use a small number of replications then the width of the confidence will be large, and we are more likely to terminate!

In a future tutorial (not yet written...) we will discuss termination criteria in more depth. For now, pick a large number of iterations and train for as long as possible.

Tip

For a rule of thumb, pick a large number of iterations to train the policy for (e.g., $10 \times |\mathcal{N}| \times \max\limits_{i\in\mathcal{N}} |\Omega_i|$)

Implementation: the training loop

The train loop of SDDP just applies the forward and backward passes iteratively, followed by a final simulation to compute the upper bound confidence interval:

function train(
    model::PolicyGraph;
    iteration_limit::Int,
    replications::Int,
    io::IO = stdout,
)
    for i in 1:iteration_limit
        println(io, "Starting iteration $(i)")
        outgoing_states, _ = forward_pass(model, io)
        backward_pass(model, outgoing_states, io)
        println(io, "| Finished iteration")
        println(io, "| | lower_bound = ", lower_bound(model))
    end
    println(io, "Termination status: iteration limit")
    μ, tσ = upper_bound(model; replications = replications)
    println(io, "Upper bound = $(μ) ± $(tσ)")
    return
end
train (generic function with 1 method)

Using our model we defined earlier, we can go:

train(model; iteration_limit = 3, replications = 100)
Starting iteration 1
| Forward Pass
| | Visiting node 1
| | | ω = 50.0
| | | x = Dict(:volume => 200.0)
| | | x′ = Dict(:volume => 0.0)
| | | C(x, u, ω) = 0.0
| | Visiting node 2
| | | ω = 0.0
| | | x = Dict(:volume => 0.0)
| | | x′ = Dict(:volume => 0.0)
| | | C(x, u, ω) = 15000.0
| | Visiting node 3
| | | ω = 0.0
| | | x = Dict(:volume => 0.0)
| | | x′ = Dict(:volume => 0.0)
| | | C(x, u, ω) = 22500.0
| Backward pass
| | Visiting node 3
| | | Skipping node because the cost-to-go is 0
| | Visiting node 2
| | | Solving φ = 0.0
| | | | V = 22500.0
| | | | dVdx′ = Dict(:volume => -150.0)
| | | Solving φ = 50.0
| | | | V = 15000.0
| | | | dVdx′ = Dict(:volume => -150.0)
| | | Solving φ = 100.0
| | | | V = 7500.0
| | | | dVdx′ = Dict(:volume => -150.0)
| | | Adding cut : 150 volume_out + cost_to_go ≥ 15000
| | Visiting node 1
| | | Solving φ = 0.0
| | | | V = 30000.0
| | | | dVdx′ = Dict(:volume => -150.0)
| | | Solving φ = 50.0
| | | | V = 22500.0
| | | | dVdx′ = Dict(:volume => -150.0)
| | | Solving φ = 100.0
| | | | V = 15000.0
| | | | dVdx′ = Dict(:volume => -150.0)
| | | Adding cut : 150 volume_out + cost_to_go ≥ 22500
| Finished iteration
| | lower_bound = 2500.0
Starting iteration 2
| Forward Pass
| | Visiting node 1
| | | ω = 0.0
| | | x = Dict(:volume => 200.0)
| | | x′ = Dict(:volume => 150.0)
| | | C(x, u, ω) = 5000.0
| | Visiting node 2
| | | ω = 50.0
| | | x = Dict(:volume => 150.0)
| | | x′ = Dict(:volume => 100.0)
| | | C(x, u, ω) = 5000.0
| | Visiting node 3
| | | ω = 100.0
| | | x = Dict(:volume => 100.0)
| | | x′ = Dict(:volume => 50.0)
| | | C(x, u, ω) = 0.0
| Backward pass
| | Visiting node 3
| | | Skipping node because the cost-to-go is 0
| | Visiting node 2
| | | Solving φ = 0.0
| | | | V = 7500.0
| | | | dVdx′ = Dict(:volume => -150.0)
| | | Solving φ = 50.0
| | | | V = 0.0
| | | | dVdx′ = Dict(:volume => -150.0)
| | | Solving φ = 100.0
| | | | V = 0.0
| | | | dVdx′ = Dict(:volume => 0.0)
| | | Adding cut : 100 volume_out + cost_to_go ≥ 12500
| | Visiting node 1
| | | Solving φ = 0.0
| | | | V = 12499.999999999998
| | | | dVdx′ = Dict(:volume => -100.0)
| | | Solving φ = 50.0
| | | | V = 7499.999999999998
| | | | dVdx′ = Dict(:volume => -100.0)
| | | Solving φ = 100.0
| | | | V = 2499.9999999999986
| | | | dVdx′ = Dict(:volume => -100.0)
| | | Adding cut : 99.99999999999999 volume_out + cost_to_go ≥ 22499.999999999996
| Finished iteration
| | lower_bound = 7499.999999999998
Starting iteration 3
| Forward Pass
| | Visiting node 1
| | | ω = 100.0
| | | x = Dict(:volume => 200.0)
| | | x′ = Dict(:volume => 200.0)
| | | C(x, u, ω) = 2499.9999999999986
| | Visiting node 2
| | | ω = 100.0
| | | x = Dict(:volume => 200.0)
| | | x′ = Dict(:volume => 150.0)
| | | C(x, u, ω) = 0.0
| | Visiting node 3
| | | ω = 100.0
| | | x = Dict(:volume => 150.0)
| | | x′ = Dict(:volume => 100.0)
| | | C(x, u, ω) = 0.0
| Backward pass
| | Visiting node 3
| | | Skipping node because the cost-to-go is 0
| | Visiting node 2
| | | Solving φ = 0.0
| | | | V = 0.0
| | | | dVdx′ = Dict(:volume => 0.0)
| | | Solving φ = 50.0
| | | | V = 0.0
| | | | dVdx′ = Dict(:volume => 0.0)
| | | Solving φ = 100.0
| | | | V = 0.0
| | | | dVdx′ = Dict(:volume => 0.0)
| | | Adding cut : cost_to_go ≥ 0
| | Visiting node 1
| | | Solving φ = 0.0
| | | | V = 7500.0
| | | | dVdx′ = Dict(:volume => -100.0)
| | | Solving φ = 50.0
| | | | V = 2500.0
| | | | dVdx′ = Dict(:volume => -100.0)
| | | Solving φ = 100.0
| | | | V = 0.0
| | | | dVdx′ = Dict(:volume => 0.0)
| | | Adding cut : 66.66666666666666 volume_out + cost_to_go ≥ 16666.666666666664
| Finished iteration
| | lower_bound = 8333.333333333332
Termination status: iteration limit
Upper bound = 9700.0 ± 865.064112763857

Success! We trained a policy for a finite horizon multistage stochastic program using stochastic dual dynamic programming.

Implementation: evaluating the policy

A final step is the ability to evaluate the policy at a given point.

function evaluate_policy(
    model::PolicyGraph;
    node::Int,
    incoming_state::Dict{Symbol,Float64},
    random_variable,
)
    the_node = model.nodes[node]
    the_node.uncertainty.parameterize(random_variable)
    for (k, v) in incoming_state
        JuMP.fix(the_node.states[k].in, v; force = true)
    end
    JuMP.optimize!(the_node.subproblem)
    return Dict(
        k => JuMP.value.(v) for
        (k, v) in JuMP.object_dictionary(the_node.subproblem)
    )
end

evaluate_policy(
    model;
    node = 1,
    incoming_state = Dict(:volume => 150.0),
    random_variable = 75,
)
Dict{Symbol, Float64} with 8 entries:
  :volume_out         => 200.0
  :demand_constraint  => 150.0
  :hydro_spill        => 0.0
  :inflow             => 75.0
  :volume_in          => 150.0
  :thermal_generation => 125.0
  :hydro_generation   => 25.0
  :cost_to_go         => 3333.33
Note

The random variable can be out-of-sample, i.e., it doesn't have to be in the vector $\Omega$ we created when defining the model! This is a notable difference to other multistage stochastic solution methods like progressive hedging or using the deterministic equivalent.

Example: infinite horizon

As promised earlier, our implementation is actually pretty general. It can solve any multistage stochastic (linear) program defined by a policy graph, including infinite horizon problems!

Here's an example, where we have extended our earlier problem with an arc from node 3 to node 2 with probability 0.5. You can interpret the 0.5 as a discount factor.

model = PolicyGraph(
    subproblem_builder;
    graph = [Dict(2 => 1.0), Dict(3 => 1.0), Dict(2 => 0.5)],
    lower_bound = 0.0,
    optimizer = HiGHS.Optimizer,
)
A policy graph with 3 nodes
Arcs:
  1 => 2 w.p. 1.0
  2 => 3 w.p. 1.0
  3 => 2 w.p. 0.5

Then, train a policy:

train(model; iteration_limit = 3, replications = 100)
Starting iteration 1
| Forward Pass
| | Visiting node 1
| | | ω = 0.0
| | | x = Dict(:volume => 200.0)
| | | x′ = Dict(:volume => 50.0)
| | | C(x, u, ω) = 0.0
| | Visiting node 2
| | | ω = 50.0
| | | x = Dict(:volume => 50.0)
| | | x′ = Dict(:volume => 0.0)
| | | C(x, u, ω) = 5000.0
| | Visiting node 3
| | | ω = 50.0
| | | x = Dict(:volume => 0.0)
| | | x′ = Dict(:volume => 0.0)
| | | C(x, u, ω) = 15000.0
| | Visiting node 2
| | | ω = 50.0
| | | x = Dict(:volume => 0.0)
| | | x′ = Dict(:volume => 0.0)
| | | C(x, u, ω) = 10000.0
| | Visiting node 3
| | | ω = 0.0
| | | x = Dict(:volume => 0.0)
| | | x′ = Dict(:volume => 0.0)
| | | C(x, u, ω) = 22500.0
| | Visiting node 2
| | | ω = 50.0
| | | x = Dict(:volume => 0.0)
| | | x′ = Dict(:volume => 0.0)
| | | C(x, u, ω) = 10000.0
| | Visiting node 3
| | | ω = 0.0
| | | x = Dict(:volume => 0.0)
| | | x′ = Dict(:volume => 0.0)
| | | C(x, u, ω) = 22500.0
| Backward pass
| | Visiting node 3
| | | Solving φ = 0.0
| | | | V = 15000.0
| | | | dVdx′ = Dict(:volume => -100.0)
| | | Solving φ = 50.0
| | | | V = 10000.0
| | | | dVdx′ = Dict(:volume => -100.0)
| | | Solving φ = 100.0
| | | | V = 5000.0
| | | | dVdx′ = Dict(:volume => -100.0)
| | | Adding cut : 49.99999999999999 volume_out + cost_to_go ≥ 4999.999999999999
| | Visiting node 2
| | | Solving φ = 0.0
| | | | V = 27500.0
| | | | dVdx′ = Dict(:volume => -150.0)
| | | Solving φ = 50.0
| | | | V = 20000.0
| | | | dVdx′ = Dict(:volume => -150.0)
| | | Solving φ = 100.0
| | | | V = 12500.0
| | | | dVdx′ = Dict(:volume => -150.0)
| | | Adding cut : 150 volume_out + cost_to_go ≥ 20000
| | Visiting node 3
| | | Solving φ = 0.0
| | | | V = 35000.0
| | | | dVdx′ = Dict(:volume => -150.0)
| | | Solving φ = 50.0
| | | | V = 27500.0
| | | | dVdx′ = Dict(:volume => -150.0)
| | | Solving φ = 100.0
| | | | V = 20000.0
| | | | dVdx′ = Dict(:volume => -150.0)
| | | Adding cut : 75 volume_out + cost_to_go ≥ 13750
| | Visiting node 2
| | | Solving φ = 0.0
| | | | V = 36250.0
| | | | dVdx′ = Dict(:volume => -150.0)
| | | Solving φ = 50.0
| | | | V = 28750.0
| | | | dVdx′ = Dict(:volume => -150.0)
| | | Solving φ = 100.0
| | | | V = 21250.0
| | | | dVdx′ = Dict(:volume => -150.0)
| | | Adding cut : 150 volume_out + cost_to_go ≥ 28749.999999999996
| | Visiting node 3
| | | Solving φ = 0.0
| | | | V = 43750.0
| | | | dVdx′ = Dict(:volume => -150.0)
| | | Solving φ = 50.0
| | | | V = 36250.0
| | | | dVdx′ = Dict(:volume => -150.0)
| | | Solving φ = 100.0
| | | | V = 28749.999999999996
| | | | dVdx′ = Dict(:volume => -150.0)
| | | Adding cut : 75 volume_out + cost_to_go ≥ 18125
| | Visiting node 2
| | | Solving φ = 0.0
| | | | V = 40625.0
| | | | dVdx′ = Dict(:volume => -150.0)
| | | Solving φ = 50.0
| | | | V = 33125.0
| | | | dVdx′ = Dict(:volume => -150.0)
| | | Solving φ = 100.0
| | | | V = 25625.0
| | | | dVdx′ = Dict(:volume => -150.0)
| | | Adding cut : 150 volume_out + cost_to_go ≥ 33125
| | Visiting node 1
| | | Solving φ = 0.0
| | | | V = 40625.0
| | | | dVdx′ = Dict(:volume => -150.0)
| | | Solving φ = 50.0
| | | | V = 33125.0
| | | | dVdx′ = Dict(:volume => -150.0)
| | | Solving φ = 100.0
| | | | V = 25625.0
| | | | dVdx′ = Dict(:volume => -150.0)
| | | Adding cut : 150 volume_out + cost_to_go ≥ 40625
| Finished iteration
| | lower_bound = 15625.0
Starting iteration 2
| Forward Pass
| | Visiting node 1
| | | ω = 100.0
| | | x = Dict(:volume => 200.0)
| | | x′ = Dict(:volume => 200.0)
| | | C(x, u, ω) = 2500.0
| | Visiting node 2
| | | ω = 100.0
| | | x = Dict(:volume => 200.0)
| | | x′ = Dict(:volume => 200.0)
| | | C(x, u, ω) = 5000.0
| | Visiting node 3
| | | ω = 0.0
| | | x = Dict(:volume => 200.0)
| | | x′ = Dict(:volume => 50.0)
| | | C(x, u, ω) = 0.0
| | Visiting node 2
| | | ω = 0.0
| | | x = Dict(:volume => 50.0)
| | | x′ = Dict(:volume => 50.0)
| | | C(x, u, ω) = 15000.0
| | Visiting node 3
| | | ω = 50.0
| | | x = Dict(:volume => 50.0)
| | | x′ = Dict(:volume => 0.0)
| | | C(x, u, ω) = 7500.0
| Backward pass
| | Visiting node 3
| | | Solving φ = 0.0
| | | | V = 48125.0
| | | | dVdx′ = Dict(:volume => -150.0)
| | | Solving φ = 50.0
| | | | V = 40625.0
| | | | dVdx′ = Dict(:volume => -150.0)
| | | Solving φ = 100.0
| | | | V = 33125.0
| | | | dVdx′ = Dict(:volume => -150.0)
| | | Adding cut : 75 volume_out + cost_to_go ≥ 20312.5
| | Visiting node 2
| | | Solving φ = 0.0
| | | | V = 35312.5
| | | | dVdx′ = Dict(:volume => -150.0)
| | | Solving φ = 50.0
| | | | V = 27812.5
| | | | dVdx′ = Dict(:volume => -150.0)
| | | Solving φ = 100.0
| | | | V = 20312.5
| | | | dVdx′ = Dict(:volume => -150.0)
| | | Adding cut : 150 volume_out + cost_to_go ≥ 35312.5
| | Visiting node 3
| | | Solving φ = 0.0
| | | | V = 42812.5
| | | | dVdx′ = Dict(:volume => -150.0)
| | | Solving φ = 50.0
| | | | V = 35312.5
| | | | dVdx′ = Dict(:volume => -150.0)
| | | Solving φ = 100.0
| | | | V = 27812.5
| | | | dVdx′ = Dict(:volume => -150.0)
| | | Adding cut : 75 volume_out + cost_to_go ≥ 21406.25
| | Visiting node 2
| | | Solving φ = 0.0
| | | | V = 17656.25
| | | | dVdx′ = Dict(:volume => -75.0)
| | | Solving φ = 50.0
| | | | V = 13906.25
| | | | dVdx′ = Dict(:volume => -75.0)
| | | Solving φ = 100.0
| | | | V = 10156.25
| | | | dVdx′ = Dict(:volume => -75.0)
| | | Adding cut : 75 volume_out + cost_to_go ≥ 28906.25
| | Visiting node 1
| | | Solving φ = 0.0
| | | | V = 26041.666666666668
| | | | dVdx′ = Dict(:volume => -100.0)
| | | Solving φ = 50.0
| | | | V = 21406.25
| | | | dVdx′ = Dict(:volume => -75.0)
| | | Solving φ = 100.0
| | | | V = 17656.25
| | | | dVdx′ = Dict(:volume => -75.0)
| | | Adding cut : 83.33333333333333 volume_out + cost_to_go ≥ 38368.055555555555
| Finished iteration
| | lower_bound = 26701.38888888889
Starting iteration 3
| Forward Pass
| | Visiting node 1
| | | ω = 0.0
| | | x = Dict(:volume => 200.0)
| | | x′ = Dict(:volume => 200.0)
| | | C(x, u, ω) = 7500.0
| | Visiting node 2
| | | ω = 100.0
| | | x = Dict(:volume => 200.0)
| | | x′ = Dict(:volume => 150.0)
| | | C(x, u, ω) = 0.0
| | Visiting node 3
| | | ω = 100.0
| | | x = Dict(:volume => 150.0)
| | | x′ = Dict(:volume => 100.0)
| | | C(x, u, ω) = 0.0
| Backward pass
| | Visiting node 3
| | | Solving φ = 0.0
| | | | V = 36041.666666666664
| | | | dVdx′ = Dict(:volume => -100.0)
| | | Solving φ = 50.0
| | | | V = 31041.666666666664
| | | | dVdx′ = Dict(:volume => -100.0)
| | | Solving φ = 100.0
| | | | V = 26041.666666666664
| | | | dVdx′ = Dict(:volume => -100.0)
| | | Adding cut : 49.99999999999999 volume_out + cost_to_go ≥ 20520.83333333333
| | Visiting node 2
| | | Solving φ = 0.0
| | | | V = 21406.25
| | | | dVdx′ = Dict(:volume => -75.0)
| | | Solving φ = 50.0
| | | | V = 18020.83333333333
| | | | dVdx′ = Dict(:volume => -49.99999999999999)
| | | Solving φ = 100.0
| | | | V = 15520.833333333328
| | | | dVdx′ = Dict(:volume => -49.99999999999999)
| | | Adding cut : 58.33333333333333 volume_out + cost_to_go ≥ 27065.97222222222
| | Visiting node 1
| | | Solving φ = 0.0
| | | | V = 26041.666666666668
| | | | dVdx′ = Dict(:volume => -100.0)
| | | Solving φ = 50.0
| | | | V = 21406.25
| | | | dVdx′ = Dict(:volume => -75.0)
| | | Solving φ = 100.0
| | | | V = 18315.97222222222
| | | | dVdx′ = Dict(:volume => -58.33333333333333)
| | | Adding cut : 77.77777777777777 volume_out + cost_to_go ≥ 37476.851851851854
| Finished iteration
| | lower_bound = 26921.296296296296
Termination status: iteration limit
Upper bound = 35033.33333333334 ± 7820.2974990575285

Success! We trained a policy for an infinite horizon multistage stochastic program using stochastic dual dynamic programming. Note how some of the forward passes are different lengths!

evaluate_policy(
    model;
    node = 3,
    incoming_state = Dict(:volume => 100.0),
    random_variable = 10.0,
)
Dict{Symbol, Float64} with 8 entries:
  :volume_out         => 0.0
  :demand_constraint  => 150.0
  :hydro_spill        => 0.0
  :inflow             => 10.0
  :volume_in          => 100.0
  :thermal_generation => 40.0
  :hydro_generation   => 110.0
  :cost_to_go         => 21406.2