Introduction to the Scipy Stack - Scientific Computing Tools for Python

Marcel Oliver


Date: February 5, 2020



Contents

Abstract:

Over the last 15 years, an enormous and increasingly well integrated collection of Python-based tools for Scientific Computing has emerged--the SciPy Stack or short SciPy [7]. This document intends to provide a quick reference guide and concise introduction to the core components of the stack. It is aimed at beginning and intermediate SciPy users and does not attempt to be a comprehensive reference manual.

1 Why Python for Scientific Computing?

Python is a full-fledged general-purpose programming language with a simple and clean syntax. It is equally well suited for expressing procedural and object oriented programming concepts, has a powerful standard library, numerous special purpose extension modules, and interfaces well with external code. Python is interpreted (strictly speaking, compiled to byte-code), ideal for interactive exploration.

As an interpreted language, Pythion cannot execute core numerical algorithms at machine speed. This is where NumPy comes in: since most large-scale computational problems can be formulated in terms of vector and matrix operations, these basic building blocks are provided with highly optimized compiled C or Fortran code and made available from within Python as array objects. Thus, any algorithm that can be formulated in terms of intrinsic operations on arrays will run at near-native speed when operating on large data sets. For those special cases where this is insufficient, Python offers mechanisms for including compiled time-critical code sections, ranging from in-lining a few lines of C to wrapping entire external libraries into Python.

On top of NumPy, a collection of several large extension libraries creates a very comprehensive, powerful environment for Scientific Computing. SciPy extends the basic array operations by providing higher-level tools which cover statistics, optimization, special functions, ODE solvers, advanced linear algebra, and more. Matplotlib is a plotting library which can create beautiful and very flexible 2D graphics as well as basic 3D plots. IPython is indispensable as an interactive shell and provides a browser-based notebook interface. Further libraries include pandas for data analysis, Mayavi for powerful 3D visualization, and SymPy for symbolic mathematics; many more specialized scientific libraries and applications are readily available on the net.

At the same time, the whole universe of general-purpose Python extensions is available to the SciPy developer. In particular, xlrd and xlwr allow reading and writing of Excel files and the csv module allows robust handling of CSV data files. There are several comprehensive GUI toolkits, libraries for network protocols, database access, and other tools which help building complex stand-alone applications.

This document focuses on the use of NumPy, the SciPy library, matplotlib, and IPython--together referred to as the SciPy Stack or simply SciPy [7]--for rapid prototyping of vectorizable numerical code. In this capacity, SciPy competes with Matlab, a proprietary development environment which has dominated Scientific Computing for many years but it is increasingly challenged by SciPy for a number of reasons:

So what are the drawbacks of SciPy? In practice, very few. Documentation is scattered and less coherent than Matlab's. When working offline, the help system is not always sufficient, especially if you do not know the precise name of a function you are looking for. This is compensated, to some extend, by lots of high quality resources on the internet. And this document tries to address some of the documentation gaps for newcomers to SciPy.

For the simplest use cases, working in Python exacts a small tribute in the form of increased complexity (name space separation, call by reference semantics). However, in relation to the associated increase in expressive power, it is a very small price to pay and handled by Python in a minimally intrusive way.

There are other considerations, both in favor of and against Python which can easily be found on the net. They are are mostly tied to specific domains of application--the above is what matters most for general purpose Scientific Computing.

2 First Steps

2.1 Software installation

Unless you have legacy code, install the stack based on Python 3. Python 2.7 is increasingly unmaintained and should not be used. On Linux, install spyder which will pull in all the software you need. (yum install python3-spyder on Redhat-based and apt-get install python3-spyder on Debian-based distributions.)

On Windows and MacOS, install the Anaconda Python distribution [4] (also available for Linux, but using distribution packages is often better).


2.2 IPython and Spyder

The Ipython shell

The Ipython shell provides a powerful command line interface to the SciPy Stack. We always invoke it as

  ipython3 --pylab
The command line option --pylab loads numpy and matplotlib into the global name space. (In general, python modules must be loaded explicitly.) It also modifies the threading model so that the shell does not block on plotting commands.

When writing scripts, you will have to load the modules you use explicitly. More on this in Section 3.1.

The Spyder interactive development environment

Spyder integrates a text editor, help browser, and interactive Ipython shell in one application. Make sure that the shell starts up with the --pylab option: Go to Tools-Preferences-Ipython console-Graphics and check the boxes “Activate support” and “Automatically load Pylab and NumPy modules”. You may also want to set “Backend” to “Automatic” to enable graphics in a separate window for zooming and panning.

2.3 Help!

2.4 Input conventions

Examples:

  In [1]: 'ab' \
     ...: 'cd'   # Explicit line continuation
  Out[1]: 'abcd'

  In [2]: min(1,
     ...:     0) # Implicit line continuation
  Out[2]: 0


2.5 Variables and simple expressions

Examples:

  In [1]: 10**2 - 1e2
  Out[1]: 0.0

  In [2]: sqrt(-1)
  Out[2]: nan

  In [3]: sqrt(-1+0j)
  Out[3]: 1j
Note: The last example shows that you have to force the use of complex numbers. This behavior is different from Matlab.

Warning: Python assignments are by reference!

All assignment in Python is by reference, not by value. When working with immutable objects such as numerical constants, Python automatically forces a copy, so the assignment-by-reference semantics is somewhat hidden. However, when working with mutable objects such as an array representing a vector or a matrix, one must force a copy when required! This behavior is a source of hard-to-discover bugs and will be discussed in detail in Section 8.

2.6 Debugging

Some commands useful for debugging:

Note: the commands above (except for del) are so-called Ipython magic functions. Should you happen to overwrite them with python function definitions, you still access them by prepending %.

2.7 Timing

To record the execution time of single code snippets or whole programs, several options are available. For most purposes, timing from the Iphython shell suffices, but you can also put timing instrumentation into any program code.

Timing code from the Ipython shell

  In [1]: %timeit sum(arange(100))
  100000 loops, best of 3: 10.8 us per loop

  In [2]: %timeit sum(arange(10000))
  10000 loops, best of 3: 33.1 us per loop

Simple CPU-time stamp

  import time
  t = time.process_time()
  # Now do some big computation
  print "Elapsed CPU Time:", time.process_time() - t
Note: use time.monotonic() instead of time.process_time() if you need wall-clock time.

Timing short code snippets

If you need a thorough analysis of small time-critical sections of code, use the timeit module:

  import timeit
  def f(x):
      # do something which needs to be timed
  t = timeit.Timer ('f(3.0)', 'from __main__ import f') 
  print ("1000 Evaluations take", t.timeit(number=1000), "s")
Note: The timeit module runs f in a controlled environment, so you have to set up the environment with the import statement. If you need to pass variables as arguments, they need to be imported, too.

3 Scripts

Except for very simple exploratory code, the normal workflow involves writing program code in separate text files and running them from the Ipython shell. We only discuss the case where the code is so simple as to be reasonably organized in a single file. You can easily modularize larger projects; consult the Python language documentation for details.

Program files should, but need not, carry the suffix .py. It is common practice to start the first line with

  #! /usr/bin/env python3
This allows direct execution of the file on Unix-like operating; on other operating systems, it is simply ignored.


3.1 Module Import

When working with script files, you are responsible for module loading. On the one hand, module import requires some explicit attention (“where is my function?!”). On the other hand, module namespaces provide sanity to complex projects and should be considered one of the strong points of working with Python.

Since Python makes it is possible to import the same function or module in different ways, it is important to standardize on some convention. We suggest the following:

Short numerical code

  from pylab import *
This import statement sets up a large set of basic, often Matlab-like functions in the current namespace. It is the script file equivalent of starting the Ipython shell with the pylab command line option as described in Section 2.2.

If a more specialized function is required, we import it by its explicit name into the current namespace. E.g., we write

  from scipy.linalg import lu
to make available the function scipy.linalg.lu as lu. This convention keeps the program code very close what you would do in the interactive shell after invoking ipython -pylab, and it also replicates Matlab to some extent.

All examples in this document use this convention!

Larger projects

When working on large projects, library modules, or programs where only a small fraction of the code is numerical, we recommend using full module path names, respectively their commonly used abbreviations. E.g.,

  import numpy as np
  import matplotlib.pyplot as plt
  import scipy.linalg
Then the common functions arange, figure, solve, all described further below, must be called as np.arange, plt.figure, and scipy.linalg.solve, respectively.

4 Arrays

array objects representing vectors, matrices, and higher-dimensional tensors are the most important building blocks for writing efficient numerical code in Python. (numpy has an alternative matrix data type more similar to the Matlab matrix model. However, its use is discouraged [6] and will not be discussed here.)


4.1 Vectors

Examples:

  In [1]: linspace(0,1,5)
  Out[1]: array([ 0.  ,  0.25,  0.5 ,  0.75,  1.  ])

  In [2]: arange(0,1,0.25)
  Out[2]: array([ 0.  ,  0.25,  0.5 ,  0.75])

  In [3]: arange(5)
  Out[3]: array([0, 1, 2, 3, 4])

  In [4]: arange(1,6)
  Out[4]: array([1, 2, 3, 4, 5])

  In [5]: arange(1,6,dtype=float)
  Out[5]: array([ 1.,  2.,  3.,  4.,  5.])

Note that arange follows the general Python convention of excluding the end point. The advantage of this convention is that the vector arange(a,b) has $b-a$ components when this difference is integer.

4.2 Matrices

A matrix $\displaystyle A = \begin{pmatrix}1 & 2 \\ 3 & 4
\end{pmatrix}$ is generated as follows.

  In [1]: A = array([[1,2],[3,4]]); A
  Out[1]: 
  array([[1, 2],
         [3, 4]])
Matrices can assembled from submatrices:
  In [2]: b = c_[5:7]; M = c_[A,b]; M
  Out[2]: 
  array([[1, 2, 5],
         [3, 4, 6]])

  In [3]: M = r_[A,b[newaxis,:]]; M
  Out[3]: 
  array([[1, 2],
         [3, 4],
         [5, 6]])

Special Matrices

There are functions to create frequently used $m \times n$ matrices. If $m=n$, only one argument is necessary.


4.3 Basic matrix arithmetic

Examples:

  In [1]: A = array([[1,2],[3,4]]); A-A.T
  Out[1]: 
  array([[ 0, -1],
         [ 1,  0]])

  In [2]: A*A.T  # Element-wise product
  Out[2]: 
  array([[ 1,  6],
         [ 6, 16]])

  In [3]: A@A.T  # Matrix product
  Out[3]: 
  array([[ 5, 11],
         [11, 25]])

Vector products

If the column vectors $v,w \in \mathbb{R}^n$ are represented by one-dimensional arrays v and w, then

Sums and products over array elements

Note that the first axis (row indices for matrices) corresponds to axis=0, the second (column indices for matrices) corresponds to axis=1, etc.

Examples

  In [1]: A = array([[1,2],[3,4]]); sum(A)                                      
  Out[1]: 10

  In [2]: sum(A,axis=0)                                                         
  Out[2]: array([4, 6])

  In [3]: sum(A,axis=1)                                                         
  Out[3]: array([3, 7])

  In [4]: cumprod(A)                                                            
  Out[4]: array([ 1,  2,  6, 24])

You can compute with Booleans where False$\equiv 0$ and True$\equiv 1$. So the following expression counts the number of nonzero elements of a matrix:

  In [4]: A = eye(3); sum(A!=0)
  Out[4]: 3

4.4 More elementary matrix operations

Examples

  In [1]: A = array([[1,2],[3,4]]); amax(A); argmax(A)
  Out[1]: 4
  Out[1]: 3

  In [2]: A.flatten()[argmax(A)]
  Out[2]: 4

  In [3]: B = 3*eye(2); argmin(B,axis=1)
  Out[3]: array([1, 0])

  In [4]: fmax(A,B)
  Out[4]: 
  array([[ 3.,  2.],
         [ 3.,  4.]])

4.5 Indexing and slicing

Python indexing starts from zero! This is often different from traditional mathematical notation, but has advantages [1].

Examples

  In [1]: A = array([[1,2],[3,4]]); A[0,0]
  Out[1]: 1

  In [2]: A[1]
  Out[2]: array([3, 4])

  In [3]: A[:,1]
  Out[3]: array([2, 4])

Ranges

Examples

  In [1]: a = arange(5); a[2:]; a[-3:]; a[::-1]
  Out[1]: array([2, 3, 4])
  Out[1]: array([2, 3, 4])
  Out[1]: array([4, 3, 2, 1, 0])

  In [2]: A = eye(4, dtype=int); A[1:3,:3] = -8; A
  Out[2]: 
  array([[ 1,  0,  0,  0],
         [-8, -8, -8,  0],
         [-8, -8, -8,  0],
         [ 0,  0,  0,  1]])

  In [3]: A[1:3,1:3] = eye(2, dtype=int)[::-1]; A
  Out[3]: 
  array([[ 1,  0,  0,  0],
         [-8,  0,  1,  0],
         [-8,  1,  0,  0],
         [ 0,  0,  0,  1]])

4.6 Array size and shape

Examples

  In [1]: A=eye(3); A.shape
  Out[1]: (3, 3)

  In [2]: A.size
  Out[2]: 9

  In [3]: len(A)
  Out[3]: 3

4.7 Reshaping arrays

Examples

  In [1]: A = arange(12); B = A.reshape(2,3,2); B
  Out[1]: 
  array([[[ 0,  1],
          [ 2,  3],
          [ 4,  5]],

         [[ 6,  7],
          [ 8,  9],
          [10, 11]]])

  In [2]: B.flatten()
  Out[2]: array([ 0,  1,  2,  3,  4,  5,  6,  7,  8,  9, 10, 11])

4.8 Index arrays

A powerful way to write short, elegant code is to use arrays of integers or Booleans to index other arrays. Though our examples only scratch the surface, many tricky index problems can be solved this way!

Examples

Select every second element from an array a:

  In [1]: a = 2**arange(8); a
  Out[1]: array([  1,   2,   4,   8,  16,  32,  64, 128])

  In [2]: i = 2*arange(4); i
  Out[2]: array([0, 2, 4, 6])

  In [3]: a[i]
  Out[3]: array([ 1,  4, 16, 64])
Note: this particular example is equivalent to a[::2], but the index array construct is much more general.

In two dimensions, index arrays work as follows. As an example, we extract the main anti-diagonal of the matrix A:

  In [4]: A = arange(16).reshape(4,4); A
  Out[4]: 
  array([[ 0,  1,  2,  3],
         [ 4,  5,  6,  7],
         [ 8,  9, 10, 11],
         [12, 13, 14, 15]])

  In [5]: i=arange(4); j=i[::-1]; A[i,j]
  Out[5]: array([ 3,  6,  9, 12])
Set all zero elements of A to $2$:
  In [6]: A = eye(3, dtype=int); A[A==0] = 2; A
  Out[6]: 
  array([[1, 2, 2],
         [2, 1, 2],
         [2, 2, 1]])

4.9 Broadcasting and generalized outer products

When performing element-wise operations on a pair of arrays with a different number of axes, Numpy will try to “broadcast” the smaller array over the additional axis or axes of the larger array provided this can be done in a compatible way. (Otherwise, an error message is raised.)

Example

  In [1]: A=ones((3,3)); b=arange(3); A*b
  Out[1]: 
  array([[ 0.,  1.,  2.],
         [ 0.,  1.,  2.],
         [ 0.,  1.,  2.]])
The detailed general rules can be found in the Numpy documentation [5]. However, you can also explicitly control how broadcasting is done by inserting the keyword newaxis into the index range specifier to indicate the axis over which this array shall be broadcast. The above example is equivalent to the following.
  In [2]: A*b[newaxis,:]
  Out[2]: 
  array([[ 0.,  1.,  2.],
         [ 0.,  1.,  2.],
         [ 0.,  1.,  2.]])
In the same way, we can have b be broadcast over columns rather than rows:
  In [3]: A*b[:,newaxis]
  Out[3]: 
  array([[ 0.,  0.,  0.],
         [ 1.,  1.,  1.],
         [ 2.,  2.,  2.]])
Finally, a more complicated example using explicit broadcasting illustrates how to compute general outer products for tensors of any shape:
  In [4]: A=array([[1,2],[3,5]]); b=array([7,11]); \
     ...: A[:,newaxis,:]*b[newaxis,:,newaxis]
  Out[4]: 
  array([[[ 7, 14],
          [11, 22]],

         [[21, 35],
          [33, 55]]])

4.10 Probability and statistics

4.11 Numerical linear algebra

The following numerical linear algebra functions are loaded into the global namespace when you import pylab.

Examples

  In [1]: A = array([[0,1],[-2,0]]); eig(A)
  Out[1]: 
  (array([ 0.+1.41421356j,  0.-1.41421356j]),
   array([[ 0.00000000-0.57735027j,  0.00000000+0.57735027j],
         [ 0.81649658+0.j        ,  0.81649658+0.j        ]]))

  In [2]: d,V = eig(A); allclose(dot(A,V), dot(V,diag(d)))
  Out[2]: True

  In [3]: svd(A)
  Out[3]: 
  (array([[ 0.,  1.],
         [ 1.,  0.]]),
   array([ 2.,  1.]),
   array([[-1., -0.],
         [ 0.,  1.]]))

  In [4]: U,s,Vh = svd(A); allclose(dot(dot(U,diag(s)),Vh),A)
  Out[4]: True

The scipy.linalg module contains many more specialized numerical linear algebra routines such as LU and Cholesky factorization, solvers for linear systems with band-matrix structure, matrix exponentials, the conjugate gradient method, and many more. To get an overview, type

  In [1]: from scipy import linalg

  In [2]: help linalg

Singular linear systems

Contrary to Matlab behavior, solve(A,b) exits with an error message when the matrix A is singular. If you need a least-square or least-norm solution, you have to use lstsq from scipy.linalg. This is illustrated in the following:

  In [1]: A = array([[1,0],[0,0]]); b=array([1,0]); c=b[::-1]

  In [2]: x = solve(A,b)
  LinAlgError: Singular matrix

  In [3]: from scipy.linalg import lstsq

  In [4]: x, res, rk, s = lstsq(A,b); x
  Out[4]: array([ 1.,  0.])
This example is an underdetermined system; any vector $x=(1,\alpha)$ would solve the system. lstsq returns the least-norm solution as well as the residual res, the effective rank of the matrix rk, and the vector of singular values s. When the system is inconsistent, the least-square solution is returned:
  In [5]: x, res, rk, s = lstsq(A,c); x
  Out[5]: array([ 0.,  0.])


5 Control structures

5.1 Branching

Conditional branching works as follows:

  if i<2:
      print("i is less than 2")
  elif i>=4:
      print("i is greater or equal to 4")
  else:
      print("i is greater or equal to 2 and less than 4")

5.2 Loops

A simple “for-loop” in Python works as follows, with the usual conventions for the range specifier.

  In [1]: for i in range(3,5):
   ...:     print(i)
  3
  4

Iterators

More generally, Python loops can iterate over many list-like objects such as arrays:

  In [1]: a = exp(0.5j*pi*arange(4))

  In [2]: for x in a:
     ...:     print(x)

  (1+0j)
  (6.12303176911e-17+1j)
  (-1+1.22460635382e-16j)
  (-1.83690953073e-16-1j)

While-loops

Here is an example of a while-loop:

  x = 5
  while x > 0:
     print(x)
     x -= 1

More flow control


5.3 Functions

Functions are defined with the keyword def.

Example

  In [1]: def square(x):
     ...:     return x*x

  In [2]: square(5)
  Out[2]: 25
Note: If a function does not take any arguments, its definition must still end with parentheses. Functions do not need to return anything; the return keyword is optional. Variables used within a function are local--changing them locally does not change a possibly globally defined variable of the same name.

Warning: References are local, not the data!

If you pass a reference to a mutable object as a function argument, you can still modify its data from within the function even though you cannot modify the reference pointer itself. However, immutable objects cannot be modified this way! This subtle difference is explained more fully in Section 8.

Universal functions

Many standard mathematical functions such as sin or abs are so-called universal functions which operate on arrays element-wise. A newly defined function f (unless it consists of nothing more than an arithmetic expression consisting only of universal functions) will not do this, but it can be “vectorized” via

  f = vectorize(f)
so that it starts acting on arrays element-wise. Note that this construct is for convenience, not for performance.

Default and named arguments

Arguments can be given a default value. The values of such optional arguments can be specified either via an ordered argument list or as an explicitly named keyword argument:

  In [1]: def root(x,q=2.0):
     ...:     return x**(1.0/q)

  In [2]: root(2), root(2,3), root(2,q=3)
  Out[2]: (1.4142135623730951, 1.2599210498948732, 1.2599210498948732)

Global variables

Use global name to declare name as a global variable.

  In [1]: def f():
     ...:     global a
     ...:     a = pi

  In [2]: a=1; f(); a
  Out[2]: 3.141592653589793

Nested functions:

In Python, it is possible to nest function definitions. Further, references to functions can be passed as references--this works just like the reference passing for data objects.

The following example is called a closure where the function defined in the inner scope has access to a local variable from the enclosing scope, but the inner function is executed only after the execution of the enclosing scope has finished.

  In [1]: def outer(xo):
     ...:     def inner(xi):
     ...:         return xo+xi
     ...:     return inner

  In [2]: f = outer(5)

  In [3]: f(3)
  Out[3]: 8

6 Graphics

6.1 Basic 2D graphs

6.2 Labels, legends, and annotation

It is possible to use TEX-style mathematical notation: enclose mathematical expressions with $-signs. Important: since TEX-commands and special characters are prefixed by a backslash, which is usually interpreted by Python as a string escape, use “raw strings” for all TEX-containing strings, e.g.,
  title(r'$f(\theta) = \sin(\theta)$')

6.3 Figure handling

6.4 3D Plots

The Matplotlib plotting package includes simple 3D graphics (for more sophisticated 3D modeling, Mayavi goes much beyond what Matplotlib can do) which has to be imported explicitly via

  from mpl_toolkits.mplot3d import Axes3D
To start a 3D plot, or more generally, a subplot, write
  ax = subplot(111, projection='3d')
Then use Consult the documentation for arguments; a simple function plotting example is given in Section 9.2 below.

7 Input and output

7.1 Console I/O

7.2 Fancy formatting

Python has a powerful format string syntax. A format string is of the form {name:format} where name is either a named argument or argument position number and format is the format specifier. Useful format specifiers are

3d
Integer of $3$ characters width
+4d Integer width $4$, print sign whether number is positive or negative
.6e Scientific notation, $6$ decimal places
8.3f Fixed point number of total width $8$ and $3$ decimal places
8.3g Choose best format for total width $8$ and $3$ significant digits

Examples

  In [1]: '|{0:+4d}|{1:4d}|{n:8d}|'.format(22,33,n=-44)
  Out[1]: '| +22|  33|     -44|'

  In [2]: '|{pi:10.3e}|{e:10.3f}|{r:10.3g}|'.format(r=sqrt(2),**locals())
  Out[2]: '| 3.142e+00|     2.718|      1.41|'

7.3 Plain text files

Reading

To open a plain text file for reading, issue

  f = open('test.txt', 'r')
To read the file contents, one often iterates over its lines as follows:
  for line in f:
      print(line)
Files are automatically closed when the file object goes out of scope, or can be explicitly closed by issuing f.close().

Writing

To open a plain text file for writing, issue

  f = open('test.txt', 'w')
This will overwrite an already existing file! If you need to append to a file, use 'a' as the mode string.
  f.write('This is line ' + str(1) + '\n')
writes the line This is line 1. Note that you have to insert explicit line termination characters '\n' to terminate a line. write accepts only strings, so you have explicitly convert other objects to strings using str. Note further that Python has an internal I/O buffer, so if you want a guaranteed update of the data on the file system, you must issue
  f.flush()
This is often useful when writing log files during long-running computations which are to be monitored via Unix tail or similar utilities. Explicitly closing the file also flushes the I/O buffer.

7.4 Saving numpy arrays as text files

7.5 Saving numpy arrays as .npz files

When you wish to save array data which only needs to be readable by other SciPy instances, it is often efficient and convenient to save them in Numpy's own compressed data format. If you need to write out more general Python objects, you should use Python's pickle module.

Writing example

  from numpy import savez
  savez('mydata.npz', A=eye(2), b=arange(2))

Reading example

  from numpy import load
  data = load('mydata.npz')
Now the two saved arrays can be accessed as data.f.A and data.f.b, respectively.

7.6 Reading CSV files

Often, CSV (“comma separated value”) files are used as a plain text data exchange format. While loadtxt can handle the simplest of CSV files, oftentimes the presence of header rows or columns, or variable numbers of columns complicate the casting of the data into a Numpy array. Consider, for example, the following tea-time order list in a file named teatime.csv:

  ,"Quantity","Price"
  "Coffee",21,1.2
  "Tea",33,0.8
The Python csv module allows robust reading of such files. It is often convenient to read in the entire file into a Python list of lists:
  import csv
  reader = csv.reader(open('teatime.csv', 'r'))
  data = [row for row in reader]
The elements of data are strings, so we might want to convert the “quantity” column into an integer array q and the “price” column into a floating point array p:
  q = array([int(row[1]) for row in data[1:len(data)]])
  p = array([float(row[2]) for row in data[1:len(data)]])
It is possible, of course, to write more sophisticated code which avoids multiple passes over the data. For adoption to different CSV coding conventions consult the csv module documentation.


8 Mutable and immutable objects

As already remarked in Section 2.5, Python variables are merely references (pointers) to a region in computer memory which stores the data. Assignments only copy the pointer, not the data itself.

Examples

  In [1]: a=ones(2); b=a; a[0]=42; print(a,b)
  [ 42.   1.] [ 42.   1.]

  In [2]: a=pi; b=a; a=42; print(a,b)
  42 3.14159265359

How can this happen? In the first example, the array ones(2) is created as a mutable object--a region in memory which contains the two floating point numbers $1.0$ and $1.0$--while a is merely a reference to this memory region. The assignment b=a copies the reference, so that any change to the data affects the array referenced by both a and b.

In the second example, after a=pi; b=a, both variables point to the object pi. However, pi is a constant and therefore immutable, so operating on a reference to it does not change the value of pi but rather creates a new reference, here to the immutable object 42. As a result, the variables a and b end up referencing different objects.

In Python, numbers, strings, and Python tuples are immutable, while arrays and Python lists and dictionaries are mutable. More generally, mutability is part of the object type specification; if in doubt, consult the specific type documentation.

Forced copy of array data

If you need a true copy of array data, you have to force it via the copy() method:

  In [3]: a=ones(2); b=a.copy(); a[0]=42; print(a,b)
  [ 42.   1.] [ 1.  1.]
Arithmetic expressions behave the way you would naturally expect:
  In [4]: a=ones(2); b=2*a; a[0]=42; print(a,b)
  [ 42.   1.] [ 2.  2.]

Mutable and immutable objects as function arguments

Data passed as a function argument can only be modified from within the function if the object type is mutable. What remains local within the scope of a function is only the reference, not the data itself. The explanation for the following example goes along the lines of the assignment example above.

  In [5]: a=ones(2); b=1

  In [6]: def f(a,b):
     ...:     a[0]=2
     ...:     b=3

  In [7]: f(a,b); print(a,b)
  [ 2.  1.] 1

9 Short self-contained examples

Figure 1: Simple plot of functions of a single variable.
\includegraphics[width=0.8\textwidth]{code/sinecosine.pdf}

9.1 Function plotting in 2D

Figure 1 shows a simple plot of scalar functions generated by the code below. The example demonstrates further how to switch to TEX-typesetting for all labels and how to annotate graphs.

  #! /usr/bin/env python3
  
  from pylab import *
  
  N = 100  # Number of plot points
  xmin = 0
  xmax = pi
  
  xx = linspace(xmin, xmax, N)
  yy = sin(xx)
  zz = cos(xx)
  
  rc('text', usetex=True) # Use TeX typesetting for all labels
  figure(figsize=(5,3))
  plot(xx, yy, 'k-',
       xx, zz, 'k:')
  xlabel('$x$')
  ylabel('$y$')
  title('Sine and Cosine on the Interval $[0,\pi]$')
  legend(('$y=\sin(x)$',
          '$y=\cos(x)$'),
         loc = 'lower left')
  annotate(r'$\sin(\frac{\pi}{4})=\cos(\frac{\pi}{4})=\frac{1}{\sqrt{2}}$',
           xy = (pi/4+0.02, 1/sqrt(2)),
           xytext = (pi/4 + 0.22, 1/sqrt(2)-0.1),
           size = 11,
           arrowprops = dict(arrowstyle="->"))
  xlim(xmin,xmax)

  savefig('sinecosine.pdf', bbox='tight')
  show()

Figure 2: A graph of a function in two variables. Note that we use a gray-scale colors to for reproduction on standard laser printers; for screen output there are visually more appealing color maps.
\includegraphics[width=0.8\textwidth]{code/plot3d.pdf}


9.2 Function plotting in 3D

The following is an example for plotting the graph of the function of two variables

$\displaystyle f(x,y) = \exp(-x^2-y^2) \,.
$

The output is shown in Figure 2
  #! /usr/bin/env python3

  from pylab import *
  from mpl_toolkits.mplot3d import Axes3D

  figure()
  ax = subplot(111, projection='3d')

  x = linspace(-3,3,40)
  xx, yy = meshgrid(x,x)
  zz = exp(-xx**2-yy**2)

  ax.plot_surface(xx, yy, zz,
                  rstride=1,
                  cstride=1,
                  cmap=cm.binary,
                  edgecolor='k',
                  linewidth=0.2)

  ax.set_xlabel(r'$x$')
  ax.set_ylabel('$y$')
  ax.set_zlabel('$z$')
  title(r'The function $z=\exp(-x^2-y^2)$')
  savefig('plot3d.pdf')
  show()

9.3 Stability of multistep methods

Stability of a numerical method for solving ordinary differential equations is of great theoretical and practical importance. A necessary condition that any method has to satisfy is zero-stability--it asserts that small perturbations in the initial data will lead to bounded perturbations of the result in the limit of small step size. Let us take the backward differentiation formula of order $6$ (BDF6) as an example. To verify zero-stability, we have to check that the roots of its first characteristic polynomial,

$\displaystyle \rho(z) = \frac{147}{60} \, z^6 -6 \, z^5 + \frac{15}{2} \, z^4
- \frac{20}{3} \, z^3 + \frac{15}{4} \, z^2 - \frac{6}{5} \, z +
\frac{1}{6} \,,
$

lie on or within the unit circle of the complex plane. For stiff equations, the region of absolute stability is also important. It is defined as the region of the complex $h\lambda $-plane in which the stability polynomial, here

$\displaystyle p(z;h\lambda) = \rho(z) - h \lambda \, z^6
$

has roots of modulus no more than $1$. For a thorough introduction to these concepts, see [3]. So the computational tasks are:
  1. Plot the roots of the first characteristic polynomial with the complex unit circle for comparison.
  2. Compute the root with the largest modulus and check if it exceeds $1$.
  3. Plot the region of absolute stability.
These tasks are accomplished by the following code; the resulting root plot is shown in Figure 3 and the region of absolute stability in Figure 4.

Figure 3: Roots of $\rho (z)$ with the complex unit circle for comparison.
\includegraphics[width=0.8\textwidth]{code/zero-stability.pdf}

Figure 4: The shaded area is the unbounded region in the complex $h\lambda $-plane where BDF6 is absolutely stable.
\includegraphics[width=0.8\textwidth]{code/abs-stability.pdf}

  #! /usr/bin/env python3

  from pylab import *

  bdf6 = array([147, -360, 450, -400, 225, -72, 10])
  R = roots(bdf6)

  c = exp(1j*linspace(0,2*pi,200))

  figure()
  plot(R.real, R.imag, 'ko',
       c.real, c.imag, 'k-')
  title('Roots of the first characteristic polynomial for BDF6')
  savefig('zero-stability.pdf', bbox='tight')
  
  # It is known that 1 is a root, so we devide it out because the
  # root condition cannot be decided numerically in this marginal case.

  z1 = array([1,-1])
  reduced_bdf6 = polydiv(bdf6, z1)[0]
  
  if max(roots(reduced_bdf6))>1.0:
      print "BDF6 is not zero-stable"
  else:
      print "BDF6 is zero-stable"
  
  # Now lets plot the region of absolute stability
  
  rhs = array([60, 0, 0, 0, 0, 0, 0])
  
  def stabroots(z):
      R = roots(bdf6 - z*rhs)
      return max(abs(R))
  stabroots = vectorize(stabroots)

  x,y = meshgrid(linspace(-10,30,350),
                 linspace(-22,22,350))
  z = stabroots(x + 1j*y)
  
  figure()
  contour(x, y, z, [1.0], colors='k')
  contourf(x, y, z, [-1.0,1.0], colors=['0.85'])
  title('Absolute stability region for BDF6')
  grid(True)
  savefig('abs-stability.pdf', bbox='tight')

  show()

9.4 Hilbert matrix

The Hilbert matrix is a standard test case for linear system solvers as it is ill-conditioned, but has a known inverse [8]. We shall test the default solver of Pylab as follows.

For each $n \times n$ Hilbert matrix $H$ where $n=1,\dots,15$ compute the solution to the linear system $H x = b$, $b = (1,\dots,1)$. Calculate the error and the condition number of the matrix and plot both in semi-logarithmic coordinates.

Solution

  #! /usr/bin/env python3

  from pylab import *
  from scipy.linalg import hilbert, invhilbert

  def test (n):
      H = hilbert(n)
      b = ones(n)
      return norm(solve(H,b) - dot(invhilbert(n),b)), cond(H)

  test = vectorize(test)

  nn = arange(1,16)
  errors, conds = test(nn)

  figure()
  semilogy(nn, errors, '-',
           nn, conds, '*')
  legend(('Error',
          'Condition Number'),
         loc = 'lower right')
  title('Hilbert matrix test case for numpy.linalg.linalg.solve')
  show()

9.5 Least square fit of a straight line

Calculate the least square fit of a straight line to the points $(x_j,y_j)$, given as two arrays x and y. Plot the points and the line.

Solution

  #! /usr/bin/env python3

  from pylab import *
  from scipy.linalg import lstsq

  def fitline (xx, yy):
      A = c_[xx, ones(xx.shape)]
      m, b = lstsq(A, yy)[0]
      return m, b

  xx = arange(21)
  yy = xx + normal(size=xx.shape)
  m, b = fitline(xx, yy)

  figure()
  plot(xx, yy, '*',
       xx, m*xx+b, '-')
  legend(('Data', 'Linear least square fit'),
         loc = 'upper left')
  show()

9.6 Q-Q plot

A quantile-quantile plot (“Q-Q plot”) is often used in statistical data analysis to visualize whether a set of measured or otherwise generated data is distributed according to another, usually assumed, probability distribution [9].

As an example, we compare samples a Poisson distribution (as synthetic “data”) to the normal distribution with the same mean and standard deviation (as the “model”); the result is shown in Figure 5.

Figure 5: Q-Q plot comparing the Poisson and the normal distribution.
\includegraphics[width=0.8\textwidth]{code/qq.pdf}

  #! /usr/bin/env python

  from pylab import *

  p = poisson(lam=10, size=4000)
  m = mean(p)
  s = std(p)
  n = normal(loc=m, scale=s, size=p.shape)

  a = m-4*s
  b = m+4*s

  figure()
  plot(sort(n), sort(p), 'o', color='0.85',
       markeredgewidth=0.2, markeredgecolor='k',)
  plot([a,b], [a,b], 'k-')
  xlim(a,b)
  ylim(a,b)
  xlabel('Normal Distribution')
  ylabel('Poisson Distribution with $\lambda=10$')
  grid(True)
  savefig('qq.pdf', bbox='tight')
  show()

Note:

In one compares an empirical distribution against an assumed model distribution, one does not need to resort to sampling the model, but can use the percentage point function (PPF), which is the inverse of the cumulative distribution function (CDF), when available. In the above example, that would amount to replacing the assignment of n by the following code:

  from scipy.stats import norm
  n = norm.ppf((0.5+arange(len(p)))/len(p), loc=m, scale=s)
(Since the resulting vector is already sorted, it also does not need explicit sorting.) When available, the PPF is the preferred way of doing the plot as avoids sampling uncertainty for the assumed model distribution.

Acknowledgments

This document is loosely based on an introduction to Octave, initially written by Hubert Selhofer [2]. Some additions were provided by Miriam Grace.

Bibliography

1
E.W. Dijkstra, Why numbering should start at zero, EWD 831, E.W. Dijkstra Archive, University of Texas at Austin, 1982.
http://www.cs.utexas.edu/users/EWD/ewd08xx/EWD831.PDF

2
H. Selhofer, with revisions by M. Oliver and T.L. Scofield, Introduction to GNU Octave.
http://math.jacobs-university.de/oliver/teaching/iub/resources/octave/octave-intro.pdf

3
E. Süli and D. Mayers, “An Introduction to Numerical Analysis,” Cambridge University Press, 2003.

4
Anaconda, accessed February 4, 2020.
https://www.anaconda.com/distribution/

5
Broadcasting, accessed September 9, 2012.
http://docs.scipy.org/doc/numpy/user/basics.broadcasting.html

6
NumPy for Matlab users, accessed February 4, 2020.
https://docs.scipy.org/doc/numpy/user/numpy-for-matlab-users.html

7
The Scipy Stack: Scientific Computing Tools for Python, accessed June 6, 2013.
http://www.scipy.org/about.html

8
“Hilbert matrix,” Wikipedia, The Free Encyclopedia, 25 March 2012, 21:02 UTC.
http://en.wikipedia.org/w/index.php?title=Hilbert_matrix&oldid=483904783

9
“Q-Q plot,” Wikipedia, The Free Encyclopedia, 10 August 2012, 23:56 UTC.
http://en.wikipedia.org/w/index.php?title=Q