Go to file
Steven G. Johnson 7800272033 add eigenproblem examples 2024-02-15 14:42:47 -05:00
notes add eigenproblem examples 2024-02-15 14:42:47 -05:00
psets Update pset4sol.ipynb: fix code comment 2023-09-01 13:49:02 -04:00
README.md link 0.30000000000000004 example 2023-05-15 14:10:06 -04:00

README.md

MIT Course 18.065/18.0651, Spring 2023

This is a repository for the course 18.065: Matrix Methods in Data Analysis, Signal Processing, and Machine Learning at MIT in Spring 2023. See also 18.065 from spring 2018 (MIT OpenCourseWare) for a previous version of this class.

Instructor: Prof. Steven G. Johnson.

Lectures: MWF1 in 2-190. Handwritten notes are posted, along with video recordings (MIT only).

Office hours (virtual): Thursdays at 4pm via Zoom.

Textbook: Linear Algebra and Learning from Data by Gilbert Strang (2019). (Additional readings will be posted in lecture summaries below.)

Resources: Piazza discussion forum, pset partners.

Grading: 50% homework, 50% final project.

Homework: Biweekly, due Fridays (2/17, 3/3, 3/17, 4/7, 4/21, 5/5) on Canvas. You may consult with other students or any other resources you want, but must write up your solutions on your own. Psets are accepted until solutions are posted (sometime Friday); extensions require permission of instructor.

Exams: None.

Final project: Due May 15 on Canvas. You can work in groups of 13 (sign up on Canvas).

  • 1-page proposal due Monday April 3 on Canvas (right after spring break), but you are encouraged to discuss it with Prof. Johnson earlier to get feedback.
  • Pick a problem involving "learning from data" (in the style of the course, but not exactly the same as what's covered in lecture), and take it further: to numerical examples, to applications, to testing one or more solution algorithms. Must include computations (using any language).
  • Final report due May 15, as an 815 page academic paper in the style template of IEEE Transactions on Pattern Analysis and Machine Intelligence.
  • Like a good academic paper, you should thoroughly reference the published literature (citing both original articles and authoritative reviews/books where appropriate [rarely web pages]), tracing the historical development of the ideas and giving the reader pointers on where to go for more information and related work and later refinements, with references cited throughout the text (enough to make it clear what references go with what results). (Note: you may re-use diagrams from other sources, but all such usage must be explicitly credited; not doing so is plagiarism.) See some previous topic areas.

What followes is a brief summary of what was covered in each lecture, along with links and suggestions for further reading. It is not a good substitute for attending lecture, but may provide a useful study guide.

Lecture 1 (Feb 6)

  • Syllabus (above) and introduction.
  • 18.065 overview diagram
  • Column space, basis, rank, rank-1 matrices, A=CR, and AB=∑(col)(row)
  • See handwritten notes and lecture video linked above.

Further reading: Textbook 1.11.3. OCW lecture 1

Lecture 2 (Feb 8)

  • Matrix multiplication by blocks and columns-times-rows. Complexity: standard algorithm for (m×p)⋅(p×n) is Θ(mnp): roughly proportional to mnp for large m,n,p, regardless of how we rearrange it into blocks. (There also exist theoretically better, but highly impractical, algorithms.)
  • Briefly reviewed the "famous four" matrix factorizations: LU, diagonalization XΛX⁻¹ or QΛQᵀ, QR, and the SVD UΣVᵀ. QR and QΛQᵀ in the columns-times-rows picture, especially QΛQᵀ (diagonalization for real A=Aᵀ) as a sum of symmetric rank-1 projections.
  • The four fundamental subspaces for an m×n matrix A of rank r, mapping "inputs" x∈ⁿ to "outputs" Ax∈ᵐ: the "input" subspaces C(Aᵀ) (row space, dimension r) and its orthogonal complement N(A) (nullspace, dimension nr); and the "output" subspaces C(A) (column space, dimension r) and its orthogonal complement N(Aᵀ) (left nullspace, dimension mr).
  • pset 1, due Friday Feb 17

Further reading: Textbook 1.31.6. OCW lecture 2. If you haven't seen matrix multiplication by blocks before, here is a nice video.

Optional Julia Tutorial: Wed Feb 8 @ 5pm via Zoom

A basic overview of the Julia programming environment for numerical computations that we will use in 18.06 for simple computational exploration. This (Zoom-based) tutorial will cover what Julia is and the basics of interaction, scalar/vector/matrix arithmetic, and plotting — we'll be using it as just a "fancy calculator" and no "real programming" will be required.

If possible, try to install Julia on your laptop beforehand using the instructions at the above link. Failing that, you can run Julia in the cloud (see instructions above).

Lecture 3 (Feb 10)

  • Orthogonal bases and unitary matrices "Q".

Choosing the right "coordinate system" (= "right basis" for linear transformations) is a key aspect of data science, in order to reveal and simplify information. The "nicest" bases are often orthonormal. (The opposite is a nearly linearly dependent "ill-conditioned" basis, which can greatly distort data.)

Orthonormal bases ⟺ QᵀQ=I, hence basis coefficients c=Qᵀx from dot products. QQᵀ is orthogonal projection onto C(Q). A square Q with orthonormal columns is known as a "orthogonal matrix" or (more generally) as a "unitary matrix": it has Qᵀ=Q⁻¹ (both its rows and columns are orthonormal). Qx preserves length ‖x‖=‖Qx‖ and dot products (angles) x⋅y=(Qx)⋅(Qy). Less obviously: any square matrix that preserves length must be unitary.

Some important examples of unitary matrices:

  • 2×2 rotation matrices
  • the identity matrix I
  • any permutation matrix P which re-orders a vector, and is simply a re-ordering of the rows/cols of I
  • Hadamard matrices: unitary matrices Hₙ/√n where Hₙ has entries of ±1 only. For n=2ᵏ they are easy to construct recursively, and are known as WalshHadamard transforms. (See also the Julia Hadamard package.)
  • discrete Haar wavelets, which are unitary after a diagonal scaling and consist of entries ±1 and 0. They are a form of "time-frequency analysis" because they reveal information about both how oscillatory a vector is ("frequency domain") and where the oscillations occur ("time domain").
  • orthonormal eigenvectors can be found for any real-symmetric ("Hermitian") matrix A=Aᵀ: A=QΛQᵀ
  • the SVD A=UΣVᵀ of any matrix A gives (arguably) the "best" orthonormal basis U for C(A) and the "best" orthonormal basis V for C(Aᵀ), which reveal a lot about A.
  • orthonormal eigenvectors can also be found for any unitary matrix! (The proof is similar to that for Hermitian matrices, but the eigenvalues |λ|=1 in this case.) Often, unitary matrices are used to describe symmetries of problems, and their eigenvectors can be thought of as a kind of "generalized Fourier transform". (All of the familar Fourier transforms, including Fourier series, sine/cosine transforms, and discrete variants thereof, can be derived in this way. For example, the symmetry of a circle gives the Fourier series, and the symmetry of a sphere gives a "spherical-harmonic transform".) For example, eigenvectors of a cyclic shift permutation give the discrete Fourier transform, which is famously computed using FFT algorithms.

Further reading: Textbook section 1.5 (orthogonality), 1.6 (eigenproblems), and 4.1 (Fourier); OCW lecture 3. The fact that preserving lengths implies unitarity is not obvious, but is proved in various textbooks; a concise summary is found here. The relationship between symmetries and Fourier-like transforms can be most generally studied through the framework of "group representation theory"; see e.g. textbooks on "group theory in physics" like Inui et al. (1996). Of course, there are whole books just on the discrete Fourier transform (DFT), just on wavelet transforms, etcetera, and you can find lots of material online at many levels of sophistication.

Lecture 4 (Feb 13)

  • Eigenproblems, diagonalization, complex inner products, and the spectral theorem.

Reviewed eigenvectors/eigenvalues Ax=λx, which make a square matrix act like a scalar λ. For an m×m matrix A, you get m eigenvalues λₖ from the roots (possibly repeated) of the characteristic polynomial det(A-λΙ), and you almost always get m independent eigenvalues xₖ (except in the very rare case of a defective matrix).

Went through the example of the 4×4 cyclic-shift permutation P from last lecture, and showed that P⁴x=x ⥰ λ⁴=1 ⥰ λ=±1,±i. We can then easily obtain the corresponding eigenvectors, and put them into the "discrete Fourier transform" matrix F.

I then claimed that the eigenvectors (properly scaled) are orthonormal, so that F is unitary, but there is a catch: we need to generalize our definition of "dot"/inner product for complex vectors and matrices. For complex vectors, the dot product x⋅y is conj(xᵀ)y (conj=complex conjugate), not xᵀy. And the length of a vector is then ‖x‖² = x⋅x = ∑ᵢ|xᵢ|², which is always ≥ 0 (and = 0 only for x=0). The "adjoint" conj(xᵀ) is sometimes denoted "xᴴ" (or x* in math, or x in physics). A unitary matrix is now more generally one for which QᴴQ=I, and we see that our Fourier matrix F is indeed unitary. And unitary matrices still preserve lengths (and inner products) with the complex inner product.

If we do a change of basis x=Bc, then Ax=λx is transformed to Cc=λc, where C=B⁻¹AB. C and A are called similar matrices, and we see that they are just the same linear operator in different bases; similar matrices always have identical eigenvalues, and eigenvectors transformed by a factor of B.

The most important change of basis for eigenproblems is changing to the basis of eigenvectors X, and showed that this gives the diagonalizaton X⁻¹AX=Λ ⟺ A = XΛX⁻¹. However, this basis may be problematic in a variety of ways if the eigenvectors are nearly dependent (X is nearly singular or "ill-conditioned", corresponding to A being nearly defective).

The nice case of diagonalization is when you have orthonormal eigenvectors X=Q, and it turns out that this arises whenever A commutes with its conjugate-transpose Aᴴ (these are called normal matrices), and give A=QΛQᴴ. Two important cases are:

  • Hermitian matrices A=Aᴴ, and especially the special case of real-symmetric matrices (real A=Aᵀ) — not only are their eigenvectors orthogonal, but their eigenvalues λ are real. In fact, if A is real-symmetric, then its eigenvectors are real too, and we can work with a purely real diagonalization A=QΛQᵀ.
  • Unitary matrices Aᴴ=A⁻¹. In this case, we can easily show that |λ|=1, and then prove orthogonality of eigenvectors from the fact that unitary matrices preserve inner products.

Further reading OCW lecture 4 and textbook section I.6.

Lecture 5 (Feb 15)

  • (Symmetric/Hermitian) positive definite ("SPD") matrices A=Aᵀ: λ > 0 ⟺ xᵀAx > 0 (for x ≠ 0) ⟺ A = BᵀB (B full column rank) ⟺ pivots > 0 (in Gaussian elimination / LU or Cholesky factorization) ⟺ ... (but it does not mean all entries of A are necessarily positive or vice versa!)
  • BᵀB matrices arise in SVDs, least-squares, statistics (covariance matrices), and many other problems.
  • Positive-definiteness of the Hessian matrix Hᵢⱼ=∂²f/∂xᵢ∂xⱼ is the key test for determining whether x∈ⁿ is a local minimum of f(x), since f(x+δ)=f(x)+∇fᵀδ + ½δᵀHδ + (higher-order). For example, H=2A for the convex quadratic "bowl" function f(x)=xᵀAx+bᵀx
  • Analogous definitions of "positive semidefinite" (> 0 replaced by ≥ 0) and negative definite/semidefinite (< 0 / ≤ 0).

Further reading OCW lecture 5 and textbook section I.7. In Julia, the isposdef function checks whether a matrix is positive definite, and does so using a Cholesky factorization (which is just Gaussian elimination speeded up 2× for SPD matrices, and fails if it encounters a negative pivot). See also these lecture slides from Stanford for more properties, examples, and applications of SPD matrices. See the matrix calculus course at MIT for a more general presentation of derivatives, gradients, and second derivatives (Hessians) of functions with vector inputs/outputs.

Lecture 6 (Feb 17)

  • "Reduced"/"compact" SVD A=UᵣΣᵣVᵣᵀ=∑σₖuₖvₖᵀ, "full" SVD A=UΣVᵀ, and "thin" SVD (what computers actually return, with zero singular values replaced by tiny roundoff errors).
  • Uᵣ and Vᵣ as the "right" bases for C(A) and C(Aᵀ). Avₖ=σₖvₖ
  • SVD and eigenvalues: U and V as eigenvectors of AAᵀ and AᵀA, respectively, and σₖ² as the positive eigenvalues.
  • Deriving the SVD from eigenvalues: the key step is showing that AAᵀ and AᵀA share the same positive eigenvalues, and λₖ=σₖ², with eigenvectors related by a factor of A. From there you can work backwards to the SVD.
  • pset 1 solutions
  • pset 2: Due March 3 at 1pm

Further reading: OCW lecture 6 and textbook section I.8. The Wikipedia SVD article.

Lecture 7 (Feb 21)

Further reading: Textbook sections I.9, I.11, III.5. OCW lecture 7 and lecture 8.

Lecture 8 (Feb 22)

Further reading: Textbook sections I.9, V.1, V.4. OCW lecture 7 and OCW lecture 20.

Lecture 9 (Feb 24)

  • pseudo-inverse A⁺=VΣ⁺Uᵀ
  • Least-squares: solve Ax≈b by minimizing ‖b-Ax‖ ⟺ solving AᵀAx̂=Aᵀb
  • 4 methods for least squares: (1) Normal equations AᵀAx̂=Aᵀb (the fastest method, but least robust to roundoff errors etc); (2) orthogonalization A=QR ⟹ Rx̂=Qᵀb (much more robust, this is essentially what A \ b does in Julia for non-square A); (3) pseudo-inverse x̂=A⁺b (pinv(A)*b in Julia; this lets you "regularize" the problem by dropping tiny singular values); (4) "ridge" or "Tikhonov" regularization (AᵀA + δ²I)⁻¹Aᵀb ⟶ x̂ as δ→0 (δ≠0 is useful to "regularize" ill-conditioned fitting problems where A has nearly dependent columns, making the solutions more robust to errors).
  • Least-squares demo

Further reading: Textbook section II.2 and OCW lecture 9. Many advanced linear-algebra texts talk about the practical aspects of roundoff errors in QR and least-squares, e.g. Numerical Linear Algebra by Trefethen and Bau (the 18.335 textbook). A nice historical review can be found in the article Gram-Schmidt orthogonalization: 100 years and more. Rarely (on a computer) explicitly form AᵀA or solve the normal equations: it turns out that this greatly exacerbates the sensitivity to numerical errors (in 18.335, you would learn that it squares the condition number). Instead, we typically use the A=QR factorization and solve Rx̂=Qᵀb. Better yet, just do A \ b (in Julia or Matlab) or the equivalent in other languages (e.g. numpy.linalg.lstsq), which will use a good algorithm. (Even professionals can get confused about this.)

Lecture 10 (Feb 27)

  • Training vs test data: VMLS slides p. 294
  • Conditioning: κ(A) = (max σ)/(min σ) is the condition number of a matrix A, and gives us a bound on the "amplification" ‖Δx‖/‖x‖ ≤ κ(a) ‖Δb‖/‖b‖ of the relative error from inputs (b) to outputs (x) when solving Ax=b (including least-squares). "Ill-conditioned problems" (κ≫1) magnify noise and other errors, and typically require some regularization (e.g. dropping smallest σ's) that trades off robustness for accuracy ‖b-Ax‖.
  • Ridge/Tikhonov/ℓ² regularization: minimize ‖b-Ax‖₂² + δ²‖x‖₂² for some penalty δ≠0 to push the solution towards smaller x. (More generally, δ²‖Dx‖₂² for some matrix D.) This gives (AᵀA+δ²I)x̂=Aᵀb, which is similar to A⁺b but replaces 1/σ with σ/(σ²+δ²). Effectively, this drops small σ's, but doesn't require an SVD and generalizes to other types of penalties. (Example: VMLS slides pg. 346.)
  • Under-determined problems: for "wide" matrices, Ax=b has many solutions (we can add any N(A) vector to a solution). A common way to pick a solution is to pick the minimum-norm solution: minimize ‖x‖₂ subject to Ax=b. (It turns out that this gives x̂=A⁺b!)

Further reading: Training/test data: VMLS section 13.2. Condition numbers: Strang exercises II.3, OCW video lecture 10, and these 18.06 notes; a more in-depth treatment can be found in e.g. Numerical Linear Algebra by Trefethen and Bau (the 18.335 textbook). Tikhonov regularization: Strang section II.2, OCW video lecture 10, VMLS section 15.3. Underdetermined minimum-norm solutions: Strang section II.2, OCW video lecture 11, UIUC Nonlinear Programming lecture notes.

Lecture 11 (Mar 1)

  • Minimum-norm solutions x̂=A⁺b=Aᵀ(AAᵀ)⁻¹: smallest ‖x‖₂ for underdetermined problems Ax=b for "wide" A.
  • Other common norms: ℓ¹ and , and sparsity with ‖x‖₁ norm (LASSO regularization).
  • Avoid AᵀA: it squares the condition number κ(AᵀA)=κ(A)²
  • GramSchmidt orthogonalization and QR factorization.

Further reading: Strang II.2, II.4.

Lecture 12 (Mar 3)

  • QR factorization: "thin" vs. "full" QR; practical considerations: pivoted QR, instability of classical GramSchmidt vs. modified GS or "GS twice" or Householder/Givens QR; the fact that many QR algorithms don't give you Q explicitly, only a fast way to multiply Qx or Qᵀy.
  • Usage of QR for least-squares problem: AᵀAx̂=Aᵀb ⥰ Rx̂=Qᵀb. (But this does not mean Ax̂=b! QQᵀ is orthogonal projection onto C(A), ≠ I in general!) This avoids squaring the condition number since κ(R)=κ(A).
  • Large-scale linear algebra: When solving m×m Ax=b for large m, we have a few options:
    • Dense linear algebra: Gaussian elimination, QR, eigenvalues, SVD, etcetera, assuming that A is just m×m numbers with no special structure. Cost is Θ(m³), memory is Θ(m²). Usually you run out of memory before running out of time! (m 10⁴ is close to filling up memory, but runs in only few minutes.)
    • Sparse-direct methods: For sparse matrices (mostly 0), only store and compute with nonzero entries. If a clever ordering is chosen for rows/cols, Gaussian elimination can often produce mostly sparse L and U factors! (This is what A \ b does in Julia if A is a sparse-matrix type.) But for very big problems even these methods can eventually run out of memory.
    • Iterative methods: start with a "guess" for x (usually x=0 or x=random), and iteratively make it closer to a solution using only A-times-vector operations (and linear combinations and dot products). Requires a fast A-times-vector, e.g. if A is sparse, low rank, a convolution, or some combination thereof. Modern methods include GMRES, BiCGSTAB(), conjugate gradient (CG), and others.
    • Randomized linear algebra: by multiplying A on the left/right by small random wide/thin matrices, carefully chosen, we can construct an approximate "sketch" of A that can be used to estimate the SVD, solutions to least-squares, etcetera, and can also accelerate iterative solvers.
    • Tricks for special cases: there are various specialized techniques for convolution/circulant matrices (via FFTs), banded matrices (linear-time methods), and low-rank updates (ShermanMorrison formula)
  • pset 2 solutions
  • pset 3 (due 3/17)

Further reading: For GramSchmidt and QR, see further reading for lecture 9. Texbook section II.1, OCW video lecture 10. Sparse-direct solvers are described in detail by the book Direct Methods for Sparse Linear Systems by Davis. Iterative methods: More advanced treatments include the book Numerical Linear Algebra by Trefethen and Bao, and surveys of algorithms can be found in the Templates books for Ax=b and Ax=λx. Some crude rules of thumb for solving linear systems (from 18.335 spring 2020).

Lecture 13 (Mar 6)

  • Continued summary of large-scale linear algebra from lecture 12, mentioning randomized algorithms (which we will cover in more detail later), such as "sketched" least-squares and randomized SVD, and also specialized algorithms for particular cases.
  • Krylov methods: defined Krylov subspaces reachable by iterative algorithms, defined a Krylov algorithm (loosely) an iterative algorithm that finds the "best" solution in the whole Krylov space (possibly approximately) on the n-th step. Gave power iteration for largest |λ| as an example of something not a Krylov method. Explained why the basis (b Ab A²b ⋯) is a poor (ill-conditioned) choice, and instead explained the Arnoldi iteration to find an orthonormal basis Qₙ by (essentially) GramSchmidt, leading to the GMRES algorithm for Ax=b.

Further reading: Arnoldi iterations and GMRES are covered in the Strang textbook section II.1, and briefly in OCW lecture 12; much more detail is found other sources (Trefethen, etc.) noted in the further reading for Lecture 12. A review of randomized linear algebra can be found in the Strang textbook sec. II.4, and also in Halko, Martinsson, and Tropp (2011). A recent paper on a variety of new randomized algorithms, e.g. for "sketched" least-square problems or to accelerate iterative algorithms like GMRES, is Nakatsukasa and Tropp (2022). A nice review of the randomized SVD can be found in a blog post by Gregory Gundersen (2019).

Lecture 14 (Mar 8)

  • Krylov wrap-up:
    • GMRES caveats: storage is Θ(mn) after n steps, and cost of orthogonalization is Θ(mn²), so in practice one is often limited to n ≲ 100. Workarounds include: "restarted GMRES", randomized "sketched GMRES", and approximate Krylov methods such as biCGSTAB(), QMR, or DQGMRES. If A is Hermitian positive-definite, however, then there is ideal Krylov method called conjugate gradient (CG) that "magically" searches the whole Krylov space using only the two most recent search directions on each iteration; CG is closely related to the "momentum" terms used in stochastic gradient descent / machine learning (covered later in 18.065).
    • GMRES convergence theory is complicated (see e.g. lecture 35 in Trefethen & Bau), but basically it converges faster if the eigenvalues are mostly "clustered" (making A more like I). You can therefore accelerate convergence by finding a preconditioner matrix M such that MA has more-clustered eigenvalues (M is a "crude inverse" of A), and then solve MAx=Mb instead of Ax=b. This can accelerate convergence by orders of magnitude, but finding a good preconditioner is a tricky problem-dependent task.
    • Krylov methods also exist for eigenproblems (e.g. Arnoldi and Jacobi-Davidson methods, or Lanczos and LOPCG for Hermitian problems), the SVD (requiring both Ax and Aᵀy operations), least-squares problems, and so on.
  • Randomized linear algebra:
    • Randomized SVD, as covered in these notes by Gregory Gundersen (2019)
    • Randomized matrix multiplication AB ≈ random sampling (col j of A)(row j of B)/pⱼ with probability pⱼ. Choosing pⱼ proportional to ‖col j of A‖⋅‖row j of B‖ minimizes the variance of this estimate! See Strang textbook section II.4 and OCW lecture 13.

Further reading: See links above, and further reading from lecture 13.

Lecture 15 (Mar 10)

  • Weighted least squares (WLS) and generalized least squares (GLS): when fitting a model with noisy measurements, we want to weight the least-squares minimization inversely with the errors (variances) in the measurements (WLS), or more generally with the inverse of the measurement correlation matrix (GLS); see textbook section V.5.
  • Reviewed correlation matrix V and its positive semidefiniteness/definiteness (textbook section V.4).
  • Framed in terms of minimizing a weighted ℓ² norm ‖b-Ax‖V⁻¹ and gave the WLS formula x̂=(AᵀV⁻¹A)⁻¹AᵀV⁻¹b = Lb.
  • This is an unbiased estimator if the measurement errors are unbiased.
  • Correlation matrix ("error bars") in the estimate x̂ is then given by W=LVLᵀ=(AᵀV⁻¹A)⁻¹ (textbook V.5).
  • GaussMarkov theorem: WLS/GLS is the unbiased linear estimator that minimizes the variance of the estimated parameters x.
    • in particular, showed that any other unbiased linear estimator gives a covariance W = W + (positive semidefinite), which results in ‖W‖≥‖W‖ in the ℓ² (induced) or Frobenius norms
    • this means that OLS squares is the best choice when the errors are unbiased, independent, and have equal variances ("homoskedastic")

Further reading: In addition to the links above, these subjects are covered in countless statistics textbooks and courses. For example, see these course notes from CMU

Lecture 16 (Mar 13)

Broad overview of optimization problems (see slides). The most general formulation is actually quite difficult to solve, so most algorithms (especially the most efficient algorithms) solve various special cases, and it is important to know what the key factors are that distinguish a particular problem. There is also something of an art to the problem formulation itself, e.g. a nondifferentiable minimax problem can be reformulated as a nicer differentiable problem with differentiable constraints.

Further reading: There are many textbooks on nonlinear optimization algorithms of various sorts, including specialized books on convex optimization, derivative-free optimization, etcetera. A useful review of topology-optimization methods can be found in Sigmund and Maute (2013).

Lecture 17 (Mar 15)

Reviewed and broadened differential calculus (18.01 and 18.02) from the perspective of 18.06, where we view a derivative f(x) as a linear operator acting on a small change in the input (dx) to give you the change in the output (df) to first order in dx ("linearized"). This viewpoint makes it easy to generalize derivatives, to scalar-valued functions of vectors where f(x) is the transposed gradient (∇f)ᵀ, to vector-valued functions of vectors where f(x) is the Jacobian matrix, and even to matrix-valued functions of matrices like f(x)=A² ⥰ df = A×dA+dA×A or f(x)=A⁻¹ ⥰ df=A⁻¹×dA×A⁻¹; in both these cases f(x) is a linear operator f(x)[dA] that takes dA in and gives df out, but cannot be written simply as (matrix)×dA.

Derivatives viewed as linear approximations have many important applications in science, machine learning, statistics, and engineering. For example, ∇f is essential for large-scale optimization of scalar-valued functions f(x), and we will see that how you compute the gradient is also critical for practical reasons. As another example, there is the multidimensional Newton algorithm for finding roots f(x)=0 of systems of nonlinear equations: At each step, you just solve a linear system of equations with the Jacobian matrix of f(x), and it converges incredibly rapidly.

Further reading: This material was presented in much greater depth in our 18.S096: Matrix Calculus course in IAP 2022 and IAP 2023. The viewpoint of derivatives as linear operators (also called Fréchet derivatives) was covered in lectures 1 and 2 of 18.S096, Newton's method was covered in lecture 4, and automatic differentiation was covered in lectures 5 and 8 — see the posted lecture materials and the further-reading links therein. This notebook has a Newton's method example demo where we solve a 2d version of the famous Thomson problem to find the equilibrium position of N repulsive "point charges" constrained to lie on a circle; more generally, a sphere or hypersphere.

Lecture 18 (Mar 17)

  • d(A⁻¹), adjoint problems, chain rules, and forward vs reverse mode (backpropagation) derivatives
  • pset 3 solutions
  • pset 4, due 4/7

Showed that d(A⁻¹)=A⁻¹ dA A⁻¹. (For example, if you have a matrix A(t) that depends on a scalar t∈, this tells you that d(A⁻¹)/dt = A⁻¹ dA/dt A⁻¹.) Applied this to computing ∇ₚf for scalar functions f(A⁻¹b) where A(p) depends on some parameters p∈ⁿ, and showed that ∂f/∂pᵢ = -(∇ₓf)ᵀA⁻¹(∂A/∂pᵢ)x. Moreover, showed that it is critically important to evaluate this expression from left to right, i.e. first computing vᵀ=-(∇ₓf)ᵀA⁻¹m, which corresponds to solving the "adjoint (transposed) equation" Aᵀv=∇ₓf. By doing so, you can compute both f and ∇ₚf at the cost of only "two solves" with matrices A and Aᵀ. This is critically important in enabling engineering/physics optimization, where you often want to optimize a property of the solution A⁻¹b of a physical model A depending on many design variables p (e.g. the distribution of materials in space).

More generally, presented the chain rule for f(g(x)) (f'(x)=g'(h(x))h'(x), where this is a composition of two linear operations, performing h' then g' — g'h' ≠ h'g'!). For functions from vectors to vectors, the chain rule is simply the product of Jacobians. Moreover, as soon as you compose 3 or more functions, it can a make a huge difference whether you multiply the Jacobians from left-to-right ("reverse-mode", or "backpropagation", or "adjoint differentiation") or right-to-left ("forward-mode"). Showed, for example, that if you have many inputs but a single output (as is common in machine learning and other types of optimization problem), that it is vastly more efficient to multiply left-to-right than right-to-left, and such "backpropagation algorithms" are a key factor in the practicality of large-scale optimization (including machine learning and neural nets).

Further reading: For the derivative of A(t)⁻¹, see Strang sec. III.1 and OCW lecture 15; for backpropagation, see Strang sec. VII.3 and OCW lecture 27. See also these slides from our matrix calculus course (lecture 4) on adjoint methods for f(A⁻¹b) and similar problems in engineering design. See the notes on adjoint methods and slides from 18.335 in spring 2021 (video) The terms "forward-mode" and "reverse-mode" differentiation are most prevalent in automatic differentiation (AD). You can find many, many articles online about backpropagation in neural networks. This video on the principles of AD in Julia by Dr. Mohamed Tarek also starts with a similar left-to-right (reverse) vs right-to-left (forward) viewpoint and goes into how it translates to Julia code, and how you define custom chain-rule steps for Julia AD.

Lecture 19 (Mar 20)

  • Line minimization and steepest descent.
  • Newton steps and the Hessian matrix H
  • Condition number κ(H) of the Hessian and convergence of steepest descent
  • Example application: iteratively solving Ax=b (for A=Aᵀ positive definite) by minimizing f(x)=xᵀAx - xᵀb. (See also the end of this example notebook.)

Further reading: Strang section VI.4; OCW lecture 21 and OCW lecture 22.

Lecture 20 (Mar 22)

Further reading: Strang section VI.4 and OCW lecture 23. Conjugate gradient as an ideal Krylov method is covered by many authors, e.g. in Trefethen and Bau lecture 38 or by Shewchuk (1994); nonlinear conjugate gradient is reviewed by Hager and Zhang (2006) and its connection to "momentum" terms is covered by e.g. Bhaya and Kaszkurewicz (2004). For accelerated gradient descent, see these lecture notes from H. Fawzi at Cambridge Univ, and this blog post by A. Wibosono at Yale. A recent article by Karimi and Vavasis (2021) presents an algorithm that blends the strengths of nonlinear conjugate gradient and accelerated gradient descent.

Lecture 21 (Mar 24)

Further reading: Strang section VI.5 and OCW lecture 25. There are many, many tutorials on this topic online. See also the links and references in the Julia notebook.

Lecture 22 (Apr 3)

In order to handle optimization problems with constraints, it's useful to first generalize the local optimality criterion ∇f₀=0 for unconstrained problems, and this leads us into Lagrangians, duality, and KKT conditions.

Started by reviewing the basic idea of Lagrange multipliers to find an extremum of one function f₀(x) and one equality constraint h₁(x)=0. We instead find an extremum of L(x,ν₁)=f₀(x)+ν₁h₁(x) over x and the Lagrange multiplier ν₁. The ν₁ partial derivative of L ensures h₁(x)=0, in which case L=f0 and the remaining derivatives extremize f0 along the constraint surface. Noted that ∇L=0 then enforces ∇f₀=0 in the direction parallel to the constraint, whereas perpendicular to the constraint ν₁ represents a "force" that prevents x from leaving the h₁(x)=0 constraint surface.

Generalized to the Lagrangian L(x,λ,ν) of the general optimization problem (the "primal" problem) with both inequality and equality constraints, following chapter 5 of the Boyd and Vandenberghe book (see below) (section 5.1.1).

Described the KKT conditions for a (local) optimum/extremum (Boyd, section 5.5.3). These are true in problems with strong duality, as pointed out by Boyd, but they are actually true in much more general conditions. For example, they hold under the "LICQ" condition in which the gradients of all the active constraints are linearly independent.

Further reading: Convex Optimization by Boyd and Vandenberghe (free book online), chapter 5. There are many sources on Lagrange multipliers (the special case of equality constraints) online that can be found by googling.

Lecture 23 (Apr 5)

Further reading: See the textbook sections III.3III.4. These slides by Stephen J. Wright at Univ. Wisc. are similar (but more in depth) to the approach from lecture. This 2011 seminar by Stephen Boyd on ADMM may also be useful, and you can find many other resources online. Many of these sources cover only equality constraints, but augmented Lagrangians can also be used for inequality constraints, e.g. as described in Birgin et al. (2007).

Lecture 24 (Apr 7)

  • Quick review of augmented Lagrangians and ADMM from last lecture. Indicator function example from section III.4.
  • CCSA interior-point algorithm
  • pset 4 solutions
  • pset 5: due 4/21

Went over very different example of a nonlinear optimization scheme, solving a fairly general inequality-constrained nonlinear-programming problem: the CCSA algorithm(s), as described by Svanberg (2002). This is a surprisingly simple algorithm (the NLopt implementation is only 300 lines of C code), but is robust and provably convergent, and illustrates a number of important ideas in optimization: optimizing an approximation to update the parameters x, guarding the approximation with trust regions and penalty terms, and optimizing via the dual function (Lagrange multipliers). Like many optimization algorithms, the general ideas are very straightforward, but getting the details right can be delicate!

Outlined the inner/outer iteration structure of CCSA, and the interesting property that it produces a sequence of feasible iterates from a feasible starting point, which means that you can stop it early and still have a feasible solution (which is very useful for many applications where 99% of optimal is fine, but feasibility is essential). It could be thought of as a type of "interior point" algorithm.

The inner optimization problem involving the approximate gᵢ functions turns out to be much easier to solve because it is convex and separable (gᵢ = a sum of 1d convex functions of each coordinate xⱼ). Convexity allows us to use strong duality to turn the problem into an equivalent "dual" optimization problem, and separability makes this dual problem trivial to formulate and solve.

Further reading: Pages 110 of Svanberg (2002) paper on CCSA algorithms — I used the "linear and separable quadratic approximation" functions gᵢ in section 5.1; as far as I can tell the other example gᵢ functions have no general advantages. (I presented a simplified form of CCSA compared to the paper, in which the per-variable scaling/trust parameters σⱼ are omitted. These can be quite useful in practice, especially if different variables have very different scalings in your problem.)

Lecture 25 (April 10)

  • Compressive sensing (CS) and ℓ¹ regularization (LASSO etc.).
  • Slides collected from various sources.

Further reading: Strang textbook, section III.5. There are many tutorials and other information on CS/LASSO/etcetera online. For example, these Rice Univ. tutorial slides (Cevher, 2019) or Princeton Slides (Cheng, 2014) are fairly accessible. For compressed sensing in MRI, see e.g. the slides by Lustig et al. and Tamir (2019) and many other sources. A hybrid of ℓ¹ (CS/LASSO) and ℓ² (ridge/Tikhonov) regularization is to use both, a combination called elastic-net regularization; see e.g. slides from Univ. Iowa (Breheny).

Lecture 26 (April 12)

Further reading: Strang section VII.1 and OCW lecture 26.

Lecture 27 (Apr 14)

  • Backpropagation for neural networks.

Further reading: Strang section VII.3 and OCW lecture 27. You can find many, many articles online about backpropagation in neural networks. For generalizing gradients to scalar-valued functions of matrices and other abstract vector spaces, what we need is an inner product; we covered this in more detail in lecture 4 of Matrix Calculus (IAP 2023). Backpropagation for neural networks is closely related to backpropagation/adjoint methods for recurrence relations (course notes), and on computational graphs (blog post); see also lecture 8 of Matrix Calculus (IAP 2023).

Lecture 28 (Apr 19)

Further reading: See section 2.1 of Moitra's Algorithmic Aspects of Machine Learning. An interesting application to image analysis can be found in this paper.

Lecture 29 (Apr 21)

Further reading: Strang textbook sections IV.2 (circulant matrices) and VII.2 (CNNs), and OCW lecture 32. See also these Stanford lecture slides and MIT lecture slides.

Lecture 30 (Apr 24)

  • Backpropagation and convolutions: differentiating with respect to convolution coefficients involves a convolution! via identity uᵀ(a⊛v)=aᵀ(Rv⊛u).
  • The discrete Fourier transform (DFT) and its inverse: roots of unity, unitarity, diagonalizing convolutions.

Further reading: Textbook section IV.1 and VII.2.

Lecture 31 (Apr 24)

Further reading: Textbook sections IV.1IV.2 and OCW lecture 31 and lecture 32. The Wikipedia FFT article (partially written by SGJ) was still not bad last I checked. Gauss and the history of the fast Fourier transform (1985) is a wonderful article on the historical development of the FFT. Duhamel & Vetterli (1990) is a classic review article. SGJ co-developed a little FFT library called FFTW.

Lecture 32 (Apr 25)

Fourier series vs. DFT: If we view the DFT as a Riemann sum approximation for a Fourier series coefficient (which turns out to be exponentially accurate for smooth periodic f(t)!), then the errors are an instance of aliasing (see e.g. the "wagon-wheel" effect). For band-limited signals where we sample at a rate > twice the bandwidth, there is no aliasing and no information loss, a result known as the Nyquist—Shannon sampling theorem; in the common case where the bandwidth is centered at ω=0, this corresponds to sampling at more than twice the highest frequency.

Further reading: For a periodic function, a Riemann sum is equivalent to a trapezoidal rule (since the 0th and Nth samples are identical), and the exponential convergence to the integral is reviewed by Trefethen and Weideman (2014); SGJ gave a simplified review for IAP (2011). The subject of aliasing, sampling, and signal processing leads to the field of digital signal processing (DSP), on which there are many books and courses. A classic textbook is Discrete-Time Signal Processing by Oppenheim and Schafer, and there are whole courses at MIT (like 6.3000 and 6.7000) on these topics.

Lecture 33 (May 1)

Further reading: See the DSP links from lecture 32.

Lecture 34 (May 3)

  • FIR filter design = polynomial fitting. Windowing & least-squares, or minimax design by Parks-McClellan and similar.
  • Changing variables H(z) = ĥ(ω) for z=exp(iω): the Z transform
  • IIR filters: definition, relation to rational functions, first-order IIR filter example, stability.

Further reading: See the DSP links from lecture 32.

Lecture 35 (May 5)

Lecture 36 (May 6)

  • Kronecker products, Hadamard matrices, and the mixed-product property (A⊗B)(C⊗D)=AC⊗BD: Kronecker products of unitary matrices are unitary.
  • Kronecker products and the fast WalshHadamard transform (FWHT): fast algorithms (FWHT, FFT, …) as sparse factorizations of dense matrices via Kronecker products, and equivalently as mono-to-multi-dimensional mappings.
  • Sylvester equations and Lyapunov equations: can be solved as ordinary matrixvector equations via Kronecker products, but this naively requires Θ(n⁶) operations, whereas exploiting the structure gives Θ(n³). Kronecker products are often a nice way to think about things but not to explicitly compute with.

Lecture 37 (May 10)

Further reading: Textbook sections VI.6VI.7 and OCW lecture 35. Using incidence matrices to identify cycles and spanning trees in graphs is also covered in 18.06 and in Strang's Introduction to Linear Algebra book (5th ed. section 10.1 or 6th ed. section 3.5), as well as in this interactive Julia notebook. A popular software package for graph and mesh partitioning is METIS. The Google PageRank algorithm is another nice application of linear algebra to graphs, in this case to rank web pages by "importance". Direct methods (e.g. Gaussian elimination) for sparse-matrix problems turn out to be all about analyzing sparsity patterns via graph theory; see e.g. the book by Timothy Davis.

Lecture 38 (May 15)

If you want to continue with "numerical computing" in any form — whether it be for data analysis, machine learning, scientific/engineering modeling, or other areas — at some point you will need to learn more about computational errors. The mathematical field that studies algorithms + approximations is called numerical analysis, covered by courses such as 18.335 at MIT.

One of the most basic sources of computational error is that computer arithmetic is generally inexact, leading to roundoff errors. The reason for this is simple: computers can only work with numbers having a finite number of digits, so they cannot even store arbitrary real numbers. Only a finite subset of the real numbers can be represented using a particular number of "bits", and the question becomes which subset to store, how arithmetic on this subset is defined, and how to analyze the errors compared to theoretical exact arithmetic on real numbers.

In floating-point arithmetic, we store both an integer coefficient and an exponent in some base: essentially, scientific notation. This allows large dynamic range and fixed relative accuracy: if fl(x) is the closest floating-point number to any real x, then |fl(x)-x| < ε|x| where ε is the machine precision. This makes error analysis much easier and makes algorithms mostly insensitive to overall scaling or units, but has the disadvantage that it requires specialized floating-point hardware to be fast. Nowadays, all general-purpose computers, and even many little computers like your cell phones, have floating-point units.

Went through some simple definitions and examples in Julia (see notebook above), illustrating the basic ideas and a few interesting tidbits. In particular, we looked at error accumulation during long calculations (e.g. summation), as well as examples of catastrophic cancellation and how it can sometimes be avoided by rearranging a calculation.

Further reading: Trefethen & Bau's Numerical Linear Algebra, lecture 13. What Every Computer Scientist Should Know About Floating Point Arithmetic (David Goldberg, ACM 1991). William Kahan, How Java's floating-point hurts everyone everywhere (2004): contains a nice discussion of floating-point myths and misconceptions. A brief but useful summary can be found in this Julia-focused floating-point overview by Prof. John Gibson. Because many programmers never learn how floating-point arithmetic actually works, there are many common myths about its behavior. (An infamous example is 0.1 + 0.2 giving 0.30000000000000004, which people are puzzled by so frequently it has led to a web site https://0.30000000000000004.com/!)