DOKK Library

Computer Science II [draft 2019/08/15]

Authors Dr. Chris Bourke

License CC-BY-SA-4.0

Plaintext
         Computer Science II

              Dr. Chris Bourke
Department of Computer Science & Engineering
      University of Nebraska—Lincoln
          Lincoln, NE 68588, USA
       http://chrisbourke.unl.edu
            cbourke@cse.unl.edu

            2019/08/15 13:02:17
               Version 0.2.0
This book is a draft covering Computer Science II topics as presented in CSCE 156
(Computer Science II) at the University of Nebraska—Lincoln.




                     This work is licensed under a Creative Commons
                     Attribution-ShareAlike 4.0 International License




                                                                                i
Contents

1 Introduction                                                                                                                                      1

2 Object Oriented Programming                                                                                                                       3
  2.1 Introduction . . . . . . . . .    .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   3
  2.2 Objects . . . . . . . . . . . .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   4
  2.3 The Four Pillars . . . . . . .    .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   4
      2.3.1 Abstraction . . . . . .     .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   4
      2.3.2 Encapsulation . . . . .     .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   4
      2.3.3 Inheritance . . . . . .     .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   4
      2.3.4 Polymorphism . . . . .      .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   4
  2.4 SOLID Principles . . . . . .      .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   4
      2.4.1 Inversion of Control . .    .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   4

3 Relational Databases                                                                                                                               5
  3.1 Introduction . . . . . . . . . . . . . .              .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .    5
  3.2 Tables . . . . . . . . . . . . . . . . .              .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .    9
      3.2.1 Creating Tables . . . . . . . . .               .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   10
      3.2.2 Primary Keys . . . . . . . . . .                .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   16
      3.2.3 Foreign Keys & Relating Tables                  .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   18
      3.2.4 Many-To-Many Relations . . .                    .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   22
      3.2.5 Other Keys . . . . . . . . . . .                .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   24
  3.3 Structured Query Language . . . . .                   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   26
      3.3.1 Creating Data . . . . . . . . . .               .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   28
      3.3.2 Retrieving Data . . . . . . . . .               .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   30
      3.3.3 Updating Data . . . . . . . . .                 .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   36
      3.3.4 Destroying Data . . . . . . . .                 .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   36
      3.3.5 Processing Data . . . . . . . . .               .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   37
  3.4 Normalization & Database Design . .                   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   47
      3.4.1 Design Example . . . . . . . . .                .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   50
  3.5 Miscellaneous . . . . . . . . . . . . .               .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   52
  3.6 Exercises . . . . . . . . . . . . . . . .             .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   53

4 List-Based Data Structures                                                              57
  4.1 Array-Based Lists . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
      4.1.1 Designing a Java Implementation . . . . . . . . . . . . . . . . . . . . . 58



                                                                                                                                                    iii
Contents

     4.2 Linked Lists . . . . . . . . . . . . . . .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   65
         4.2.1 Designing a Java Implementation        .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   71
         4.2.2 Variations . . . . . . . . . . . . .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   74
     4.3 Stacks & Queues . . . . . . . . . . . .      .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   77
         4.3.1 Stacks . . . . . . . . . . . . . . .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   77
         4.3.2 Queues . . . . . . . . . . . . . . .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   82
         4.3.3 Variations . . . . . . . . . . . . .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   86

5 Algorithm Analysis                                                                                                                    91
  5.1 Introduction . . . . . . . . . . . . . . . . . .            . . . . . . .               .   .   .   .   .   .   .   .   .   .   . 91
      5.1.1 Example: Computing a Sum . . . . . .                  . . . . . . .               .   .   .   .   .   .   .   .   .   .   . 97
      5.1.2 Example: Computing a Mode . . . . .                   . . . . . . .               .   .   .   .   .   .   .   .   .   .   . 99
  5.2 Pseudocode . . . . . . . . . . . . . . . . . .              . . . . . . .               .   .   .   .   .   .   .   .   .   .   102
  5.3 Analysis . . . . . . . . . . . . . . . . . . . .            . . . . . . .               .   .   .   .   .   .   .   .   .   .   105
  5.4 Asymptotics . . . . . . . . . . . . . . . . . .             . . . . . . .               .   .   .   .   .   .   .   .   .   .   109
      5.4.1 Big-O Analysis . . . . . . . . . . . . .              . . . . . . .               .   .   .   .   .   .   .   .   .   .   109
      5.4.2 Other Notations . . . . . . . . . . . . .             . . . . . . .               .   .   .   .   .   .   .   .   .   .   111
      5.4.3 Observations . . . . . . . . . . . . . .              . . . . . . .               .   .   .   .   .   .   .   .   .   .   113
      5.4.4 Limit Method . . . . . . . . . . . . . .              . . . . . . .               .   .   .   .   .   .   .   .   .   .   116
  5.5 Examples . . . . . . . . . . . . . . . . . . .              . . . . . . .               .   .   .   .   .   .   .   .   .   .   118
      5.5.1 Linear Search . . . . . . . . . . . . . .             . . . . . . .               .   .   .   .   .   .   .   .   .   .   118
      5.5.2 Set Operation: Symmetric Difference .                 . . . . . . .               .   .   .   .   .   .   .   .   .   .   120
      5.5.3 Euclid’s GCD Algorithm . . . . . . . .                . . . . . . .               .   .   .   .   .   .   .   .   .   .   121
      5.5.4 Selection Sort . . . . . . . . . . . . . .            . . . . . . .               .   .   .   .   .   .   .   .   .   .   123
  5.6 Other Considerations . . . . . . . . . . . . .              . . . . . . .               .   .   .   .   .   .   .   .   .   .   124
      5.6.1 Importance of Input Size . . . . . . . .              . . . . . . .               .   .   .   .   .   .   .   .   .   .   124
      5.6.2 Control Structures are Not Elementary                 Operations                  .   .   .   .   .   .   .   .   .   .   127
      5.6.3 Average Case Analysis . . . . . . . . .               . . . . . . .               .   .   .   .   .   .   .   .   .   .   128
      5.6.4 Amortized Analysis . . . . . . . . . . .              . . . . . . .               .   .   .   .   .   .   .   .   .   .   129
  5.7 Analysis of Recursive Algorithms . . . . . .                . . . . . . .               .   .   .   .   .   .   .   .   .   .   130
      5.7.1 The Master Theorem . . . . . . . . . .                . . . . . . .               .   .   .   .   .   .   .   .   .   .   131
  5.8 Exercises . . . . . . . . . . . . . . . . . . . .           . . . . . . .               .   .   .   .   .   .   .   .   .   .   134

6 Trees                                                                                                                               137
  6.1 Introduction . . . . . . . . . . . . . .    .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   137
  6.2 Definitions & Terminology . . . . . .       .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   137
  6.3 Implementation . . . . . . . . . . . .      .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   146
  6.4 Tree Traversal . . . . . . . . . . . . .    .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   147
      6.4.1 Preorder Traversal . . . . . . .      .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   148
      6.4.2 Inorder Traversal . . . . . . . .     .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   150
      6.4.3 Postorder Traversal . . . . . . .     .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   153
      6.4.4 Tree Walk Traversal . . . . . .       .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   157
      6.4.5 Breadth-First Search Traversal        .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   159



iv
                                                                                                                                          Contents

  6.5 Binary Search Trees . .     .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   161
      6.5.1 Retrieval . . . . .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   162
      6.5.2 Insertion . . . . .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   163
      6.5.3 Removal . . . . .     .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   165
      6.5.4 In Practice . . .     .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   167
  6.6 Heaps . . . . . . . . .     .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   169
      6.6.1 Operations . . .      .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   170
      6.6.2 Implementations       .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   175
      6.6.3 Variations . . . .    .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   179
      6.6.4 Applications . . .    .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   180
  6.7 Exercises . . . . . . . .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   183

Appendix                                                                         185
  1 Author-Book Database SQL . . . . . . . . . . . . . . . . . . . . . . . . . . 185

Glossary                                                                                                                                              187

Acronyms                                                                                                                                              191


Index                                                                                                                                                 195


References                                                                                                                                            195




                                                                                                                                                        v
List of Algorithms
 1    Insert-At-Head Linked List Operation         .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   66
 2    Insert Between Two Nodes Operation           .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   67
 3    Index-Based Retrieval Operation . . .        .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   69
 4    Key-Based Delete Operation . . . . .         .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   71

 5    Computing the Mean . . . . . . . . .         .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   103
 6    Computing the Mode . . . . . . . . .         .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   104
 7    Trivial Sorting (Bad Pseudocode) . . .       .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   104
 8    Trivially Finding the Minimal Element        .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   105
 9    Finding the Minimal Element . . . . .        .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   105
 10   Linear Search . . . . . . . . . . . . . .    .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   119
 11   Symmetric Difference of Two Sets . .         .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   120
 12   Euclid’s GCD Algorithm . . . . . . . .       .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   121
 13   Selection Sort . . . . . . . . . . . . . .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   123
 14   Sieve of Eratosthenes . . . . . . . . .      .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   125
 15   Fibonacci(n) . . . . . . . . . . . . . .     .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   131
 16   Binary Search – Recursive . . . . . . .      .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   133
 17   Merge Sort . . . . . . . . . . . . . . .     .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   133

 18   Stack-based Preorder Tree Traversal . . . . . . . . . . . .                              .   .   .   .   .   .   .   .   .   .   149
 19   preOrderTraversal(u): Recursive Preorder Tree Traversal .                                .   .   .   .   .   .   .   .   .   .   150
 20   Stack-based Inorder Tree Traversal . . . . . . . . . . . . .                             .   .   .   .   .   .   .   .   .   .   151
 21   inOrderTraversal(u): Recursive Inorder Tree Traversal . .                                .   .   .   .   .   .   .   .   .   .   153
 22   Stack-based Postorder Tree Traversal . . . . . . . . . . . .                             .   .   .   .   .   .   .   .   .   .   154
 23   postOrderTraversal(u): Recursive Postorder Tree Traversal                                .   .   .   .   .   .   .   .   .   .   157
 24   Tree Walk based Tree Traversal . . . . . . . . . . . . . . .                             .   .   .   .   .   .   .   .   .   .   160
 25   Queue-based BFS Tree Traversal . . . . . . . . . . . . . .                               .   .   .   .   .   .   .   .   .   .   161
 26   Search algorithm for a binary search tree . . . . . . . . . .                            .   .   .   .   .   .   .   .   .   .   163
 27   Finding the maximum key value in a node’s left subtree. .                                .   .   .   .   .   .   .   .   .   .   167
 28   Heapify . . . . . . . . . . . . . . . . . . . . . . . . . . . .                          .   .   .   .   .   .   .   .   .   .   171
 29   Find Next Open Spot - Numerical Technique . . . . . . .                                  .   .   .   .   .   .   .   .   .   .   179
 30   Heap Sort . . . . . . . . . . . . . . . . . . . . . . . . . . .                          .   .   .   .   .   .   .   .   .   .   181




                                                                                                                                           vii
List of Code Samples
 3.1 A simple table definition. We use this as an illustrative example, it is                                             not
     necessarily a well-designed table. . . . . . . . . . . . . . . . . . . . .                                           . . . . 10
 3.2 A many-to-many Author/Book table design . . . . . . . . . . . . . .                                                  . . . . 23
 3.3 Full Author/Book table database . . . . . . . . . . . . . . . . . . . .                                              . . . . 27

 4.1 Parameterized Array-Based List in Java . . . . . . . . . . . . . . . . . . . . 88
 4.2 A linked list node Java implementation. Getter and setter methods have
     been omitted for readability. A convenience method to determine if a node
     has a next element is included. This implementation uses null as its
     terminating value. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89

 5.1   Summing a collection of integers   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   . 94
 5.2   Summation Algorithm 1 . . . . .    .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   . 97
 5.3   Summation Algorithm 2 . . . . .    .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   . 97
 5.4   Summation Algorithm 3 . . . . .    .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   . 97
 5.5   Mode Finding Algorithm 1 . . .     .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   . 99
 5.6   Mode Finding Algorithm 2 . . .     .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   100
 5.7   Mode Finding Algorithm 3 . . .     .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   101
 5.8   Naive Exponentiation . . . . . .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   126
 5.9   Computing an Average . . . . .     .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   127




                                                                                                                                        ix
List of Figures
 3.1    CSV Formatted Data . . . . . . . . . . . . . . . . . . . . . .    .   .   .   .   .   .   .   .    6
 3.2    JSON and XML data representation examples . . . . . . . .         .   .   .   .   .   .   .   .    8
 3.3    Parent Child Table Relation . . . . . . . . . . . . . . . . . .   .   .   .   .   .   .   .   .   20
 3.4    Entity-Relation Diagram . . . . . . . . . . . . . . . . . . . .   .   .   .   .   .   .   .   .   21
 3.5    A many-to-many relationship using a join table. . . . . . . .     .   .   .   .   .   .   .   .   24
 3.6    Author/Book Many-to-Many Data . . . . . . . . . . . . . .         .   .   .   .   .   .   .   .   25
 3.7    Final Author-Book Database . . . . . . . . . . . . . . . . . .    .   .   .   .   .   .   .   .   28
 3.8    A 3-D cube projected onto a 2-D surface producing a square.       .   .   .   .   .   .   .   .   38
 3.9    Enrollment Database . . . . . . . . . . . . . . . . . . . . . .   .   .   .   .   .   .   .   .   52
 3.10   Enrollment Data . . . . . . . . . . . . . . . . . . . . . . . .   .   .   .   .   .   .   .   .   53
 3.11   Simple Video Game Database . . . . . . . . . . . . . . . . .      .   .   .   .   .   .   .   .   54

 4.1 A simple linked list containing 3 nodes. . . . . . . . . . . . . . . . . . . .                   . 66
 4.2 Insert-at-head Operation in a Linked List. We wish to insert a new
      element, 42 at the head of the list. . . . . . . . . . . . . . . . . . . . . . .                . 67
 4.3 Inserting Between Two Nodes in a Linked List. Here, we wish to insert a
      new element 42 between the given two nodes containing 8 and 6. . . . . .                        . 68
 4.4 Delete Operation in a Linked List . . . . . . . . . . . . . . . . . . . . . .                    . 70
 4.5 Key-Based Find and Remove Operation. We wish to remove the first node
      we find containing 42. . . . . . . . . . . . . . . . . . . . . . . . . . . . . .                .   72
 4.6 A Doubly Linked List Example . . . . . . . . . . . . . . . . . . . . . . .                       .   75
 4.7 A Circularly Linked List Example . . . . . . . . . . . . . . . . . . . . . .                     .   76
 4.8 An Unrolled Linked List Example . . . . . . . . . . . . . . . . . . . . . .                      .   76
 4.9 A stack holding integer elements. Push and pop operations are depicted
      as happening at the “top” of the stack. In actuality, a stack stored in a
      computer’s memory is not really oriented but this visualization is consistent
      with a physical stack growing “upwards.” . . . . . . . . . . . . . . . . . .                    . 79
 4.10 An example of a queue. Elements are enqueued at the end of the queue
      and dequeued from the front of the queue. . . . . . . . . . . . . . . . . .                     . 82
 4.11 Array-Based Queue . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .                   . 85
 4.12 Producer Consumer Pattern . . . . . . . . . . . . . . . . . . . . . . . . .                     . 86

 5.1 Quadratic Regression of Index-Based Linked List Performance . . . . . . . 96
 5.2 Plot of two functions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108
 5.3 Expected number of comparisons for various success probabilities p. . . . 129

 6.1 An undirected graph with labeled vertices. . . . . . . . . . . . . . . . . . 138



                                                                                                          xi
List of Figures

      6.2 Several examples of trees. It doesn’t matter how the tree is depicted or
           organized, only that it is acyclic. The final example, 6.2(d) represents a
           disconnected tree, called a forest. . . . . . . . . . . . . . . . . . . . . . . 140
      6.3 Several possible rooted orientations for the tree from Figure 6.2(a). . . . 141
      6.4 A Tree Node’s Relations. . . . . . . . . . . . . . . . . . . . . . . . . . . . 143
      6.5 A Binary Tree . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 143
      6.6 A complete tree of depth d = 3 which has 1 + 2 + 4 + 8 = 15 nodes. . . . 144
      6.7 A summation of nodes at each level of a complete binary tree up to depth d.145
      6.8 A small binary tree. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 148
      6.9 A walkthrough of a preorder traversal on the tree from Figure 6.5. . . . . 149
      6.10 A walkthrough of a inorder traversal on the tree from Figure 6.5. . . . . . 152
      6.11 A walkthrough of a postorder traversal on the tree from Figure 6.5. . . . 156
      6.12 A tree walk on the tree from Figure 6.8. . . . . . . . . . . . . . . . . . . 158
      6.13 The three general cases of when to process a node in a tree walk. . . . . 158
      6.14 A Breadth First Search Example . . . . . . . . . . . . . . . . . . . . . . 159
      6.15 A Binary Search Tree . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 162
      6.16 Various Search Examples on a Binary Search Tree . . . . . . . . . . . . . 164
      6.17 Binary Search Tree Insertion Operation . . . . . . . . . . . . . . . . . . . 165
      6.18 Binary Search Tree Deletion Operation Exampless . . . . . . . . . . . . . 168
      6.19 A degenerate binary search tree. . . . . . . . . . . . . . . . . . . . . . . . 168
      6.20 A min-heap . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 169
      6.21 A Max-heap . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 170
      6.22 An Invalid Max-heap . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 171
      6.23 Insertion and Heapification . . . . . . . . . . . . . . . . . . . . . . . . . . 172
      6.24 Another Invalid Heap . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 173
      6.25 Removal of the root element (getMax) and Heapification . . . . . . . . . 174
      6.26 Heap Node’s Index Relations . . . . . . . . . . . . . . . . . . . . . . . . . 176
      6.27 An array implementation of the heap from Figure 6.21 along with the
           generalized parent, left, and right child relations. . . . . . . . . . . . . . . 176
      6.28 Tree-based Heap Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . 178




xii
1 Introduction
    “Smart data structures and dumb code works a lot better than the other way
    around”
                                                                  — Eric S. Raymond [8]

To Come.




                                                                                 1
2 Object Oriented Programming




2.1 Introduction




To Come.
Alan Kay (Pronounced oooo-p): ”the main thing about doing oop work or any kind of
programming work is that there has to be some exquisite blend between beauty and
practicality. There is no reason to sacrifice either one of those and people who are willing
to sacrifice either one of those I think, don’t really get what computing is all about.
(OOPSLA 1997); Alan Kay created smalltalk and coined ”OOP”
As things become more complex, architecture outstrips material.
maybe; General Responsibility Assignment Software Patterns (GRASP)



                                                                                          3
2 Object Oriented Programming


2.2 Objects

2.3 The Four Pillars

2.3.1 Abstraction

2.3.2 Encapsulation

2.3.3 Inheritance

2.3.4 Polymorphism

2.4 SOLID Principles

2.4.1 Inversion of Control

solId idea: Inversion of Control: 3 GPS data provider (Android provider using Location +
float for lat/lon + certain method names, 2nd that is more accurate, but uses doubles, no
location object, etc.: problems with adapting to ”set” a marker on a map; then introduce
a 3rd provider to show that even more code needs to change; then invert the control by
defining a single interface and adapters for each provider




4
3 Relational Databases

3.1 Introduction

Consider the data in Table 3.1 for a course enrollment system.

  Course       Course Title             Student         NUID       Email
                                        waits, tom      10001949   tomwaits@hotmail.com
  CSCE 156     Intro to CSII            Lou Reed        10001942   reed@gmail.com
  CSCE 156     Introduction to CS2      Tom Waits       10001949   twaits@email.com
  CSCE 230     Computer Hardware        Student, J.     12345678   jstudent@unl.edu
  CSCE 156     Intro to CSII            Student, John   12345678   jstudent@unl.edu
  MATH 479     Wonton Burrito Meals     Philip J. Fry   10002000   fry@unl.edu
  CSCE 235     Discrete Math            Student, John   12345678   jstudent@unl.edu
  CSCE 235     Discrete Math            Student, John   12345678   jstudent@cse.unl.edu
  CSCE 235     Discrete Math            Tom Waits       10001942   twaits@email.com
  CSCE 235     Discrete Math            Lou Reed        10001942   reed@gmail.com
  NONE         null                     Lou Reed        10001942   reed@gmail.com

                              Table 3.1: Course Enrollment Data

The data in this table is essentially a flat file. Though we’ve presented it in a nicely
formatted table, this data is essentially no different from a Comma-Separated Value (CSV)
file in which each line represents a record and individual column values are delimited with
a comma. The same data as a CSV file is presented in Figure 3.1. Flat file representations
of data are called such because they take related pieces of data, such as which courses a
student is enrolled, and flatten them into a single table representation. Even ancillary
data that have a completely different relationship (such as emails) is included in the
“flattened” representation.
Flat file representations can lead to a lot of problems, limitations and can lead to data
anomalies. To motivate the use of a proper database system, we identify several of these
here.
    • There is no semantic meaning to the data fields. Though the first row contains
      headers of what each column represents, there is no way to enforce those associations.
      Since this data format is merely a plaintext file, we could accidentally swap columns,



                                                                                          5
3 Relational Databases


    Course,Course Title,Student,NUID,Email
    ,,waits,tom,10001949,tomwaits@hotmail.com
    CSCE 156,Intro to CSII,Lou Reed,10001942,reed@gmail.com
    CSCE 156,Introduction to CS2,Tom Waits,10001949,twaits@email.com
    CSCE 230,Computer Hardware,Student,J.,12345678,jstudent@unl.edu
    CSCE 156,Intro to CSII,Student,John,12345678,jstudent@unl.edu
    MATH 479,Wonton Burrito Meals,Philip J. Fry,10002000,fry@unl.edu
    CSCE 235,Discrete Math,Student,John,12345678,jstudent@unl.edu
    CSCE 235,Discrete Math,Student,John,12345678,jstudent@cse.unl.edu
    CSCE 235,Discrete Math,Tom Waits,10001942,twaits@email.com
    CSCE 235,Discrete Math,Lou Reed,10001942,reed@gmail.com
    NONE,null,Lou Reed,10001942,reed@gmail.com


                            Figure 3.1: CSV Formatted Data


      omit columns or provide more tokens than expected. Alternatively, some data
      representation formats are designed for Electronic Data Interchange (EDI). In order
      to transfer data between two different systems (that may be on different hardware,
      use different data models or are written in different languages) data needs to be
      translated (a process called serialization) into a universal format in which each data
      field is semantically marked up to indicate what it represents. Two examples can be
      found in Figure 3.2 which contains samples of the same enrollment data represented
      as JavaScript Object Notation (JSON) and Extensible Markup Language (XML)
      respectively.
    • There is a lot of repetition of data. Every record that models a student enrolled
      in a course repeats all the student’s data including their name, ID, email, etc.
      Likewise, the course information is also repeated over every enrollment record.
    • Repeated data can lead to different and conflicting representations and for-
      matting problems. In some of the student records the name is similar, but
      represented differently, some are last name first, others are first name first, others
      have a middle initial or only have an initial for the first name, one doesn’t use
      capitalization, etc.
    • There is missing or incomplete data. The first and last record do not have any
      course data and they represent this missing data inconsistently: with empty string
      values in the first record, a placeholder value (“NONE”), and a null value. These
      may be records that once held data but that was changed. However, so as not to
      lose the (still relevant) email data, they were modified.
    • There is inconsistent data. The third from last record for example has a different
      NUID for Tom Waits as it does from the other records with a similar name. This
      NUID is also associated with Lou Reed in the final two records. Without any rules
      to enforce consistency, there is no data integrity and incorrect data like this is



6
                                                                         3.1 Introduction

      allowed to occur.
   • There are organizational and efficiency issues with processing a single data
     file. Any aggregate reports, such as producing a roster for a particular class or a
     schedule for a single student, would require processing every record in the entire
     file. Similarly, updating a single record becomes extremely difficult. For example,
     if one student change their email, we have to find and replace every instance with
     several contingencies (dealing with multiple email records) and side effects to deal
     with.
   • Keeping data in a single file also has concurrency Issues. If we want to design
     a system to be multi-user or even simply multithreaded or have several programs
     access and process our data then each process or thread may have to do so by
     placing a file lock on the data file. This precludes the possibility of concurrent and
     parallel processing, severely limiting a system’s efficiency and essentially making it
     a single user system.

The solution to (most) of these problems is to use a proper Relational Database Manage-
ment System (RDMS). A relational database stores data in tables. Tables delineate data
in columns and rows. Each column holds a specific type of data (integers, strings, etc.)
that define fields of records. Each row in a table corresponds to a single record.
Moreover, each row can be uniquely identified using a unique primary key which can be
automatically managed by the database. Data between tables can be related to each
other using foreign keys. A database enables you to define rules and constraints to enforce
data integrity. Each column has a particular type and can be made to be required with
invalid values rejected. Separating data into different tables and relating data between
tables reduces the failure points, reduces redundancy and minimizes the potential for
data integrity. Databases are sophisticated programs that structure data and provide
better organization. They provide features for concurrency, authentication, security, and
more that address the problems identified previously.
The key features that an RDMS offers can be summed up with the Atomicity, Concurrency,
Isolation, Durability (ACID) principles. In a database, the basic unit of work is known
as a transaction. A transaction may consist of one or more queries which are individual
requests for data (retrieving current records) or requests for operations on data in
a database (creating new records, updating or deleting current records). Databases
treat transactions as an all-or-nothing operation that is either committed (changes are
permanently made to the data in a database) or rolled back (there was a problem with
the query or a potential data integrity issue that would make the operation(s) invalid).
The ACID principles ensure that every transaction has the following four properties.

   • Atomicity: transactions must be an all-or-nothing operation. Suppose a trans-
     action consists of multiple queries that modify the data and that the third query
     would violate the constraints (rules) of the database. The transaction would fail
     and rollback as if the first two queries never happened. Atomic transactions are



                                                                                         7
3 Relational Databases


     1   [
     2       {
     3            "Course": "",
     4            "Course Title": "",
     5            "Student": "waits",
     6            "NUID": "tom",
     7            "Email": 10001949
     8       },
     9       {
    10            "Course": "CSCE 156",
    11            "Course Title": "Intro to CSII",
    12            "Student": "Lou Reed",
    13            "NUID": 10001942,
    14            "Email": "reed@gmail.com"
    15       },
    16       ...
    17       {
    18          "Course": "NONE",
    19          "Course Title": null,
    20          "Student": "Lou Reed",
    21          "NUID": 10001942,
    22          "Email": "reed@gmail.com"
    23       }
    24   ]


     1   <?xml version="1.0" encoding="UTF-8" ?>
     2   <enrollments>
     3     <enrollment>
     4       <Course></Course>
     5       <Student>waits</Student>
     6       <NUID>tom</NUID>
     7       <Email>10001949</Email>
     8       <Course_Title></Course_Title>
     9     </enrollment>
    10     <enrollment>
    11       <Course>CSCE 156</Course>
    12       <Student>Lou Reed</Student>
    13       <NUID>10001942</NUID>
    14       <Email>reed@gmail.com</Email>
    15       <Course_Title>Intro to CSII</Course_Title>
    16     </enrollment>
    17     ...
    18     <enrollment>
    19       <Course>NONE</Course>
    20       <Student>Lou Reed</Student>
    21       <NUID>10001942</NUID>
    22       <Email>reed@gmail.com</Email>
    23       <Course_Title/>
    24     </enrollment>
    25   </enrollments>


                       Figure 3.2: JSON and XML data representation examples


8
                                                                                3.2 Tables

      not divisible or decomposable.
   • Consistency: all transactions will retain a state of consistency so that all con-
     straints and requirements of the database will be preserved and all data will
     conform to them before and after the transaction is committed. For efficiency, a
     database may temporarily violate constraints during a transaction, but consistency
     is guaranteed afterwards.
   • Isolation: no transaction interferes with or is even aware of another. This ensures
     that we have concurrency and that different threads, programs or users will not
     interfere with each other.
   • Durability: once committed, a transaction remains so, preserving the constraints,
     rules and integrity of the data.

The relational data model was originally developed by Edgar Codd [5] in 1974. Today there
are many relational database systems ranging from Free and Open Source Software (FOSS)
(SQLite, PostgreSQL, MariaDB) to commercial software (SQLServer, DB2, Oracle)
costing hundreds or even millions of dollars. The Structured Query Language (SQL) is
the defacto standard for interacting with these databases (designed by Chamberlin &
Boyce [4] at IBM). Each one of these systems has their own dialect of SQL (PostgreSQL
uses Psql, SQLServer uses Transact-SQL) that offers vendor-specific functionality and
features. However, most of them do support the core features of the language which we’ll
focus on.


3.2 Tables

The first step to creating a database is to define tables that will hold data. When creating
tables, you specify the table’s name as well as its columns and the type of data each
column represents. Many tools exist that allow you to create a database using a graphical
user interface or even to programmatically generate a database using existing data. We’ll
focus on using basic SQL to create our tables. Typically, scripts to create tables are
stored in Document Description Language (DDL) files which are simply plaintext files
containing the commands to generate a database. You can also add comments to your
SQL scripts using two hyphens, -- or sometimes the hash symbol # .




                                                                                          9
3 Relational Databases


     1   -- This table will be used to hold data on books
     2   -- ISBN = International Standard Book Number
     3   -- dewey corresponds to the Dewey Decimal number
     4   create table Book (
     5     bookId integer primary key auto_increment not null,
     6     title varchar(255) not null,
     7     author varchar(255),
     8     isbn   varchar(255) not null default '',
     9     dewey float,
 10        numCopies integer default 0
 11      );

Code Sample 3.1: A simple table definition. We use this as an illustrative example, it is
                 not necessarily a well-designed table.


3.2.1 Creating Tables

To create a table, we use the following basic syntax.

     1   create table TableName (
     2     columnA columnType [options],
     3     columnB columnType [options],
     4     ...
     5   );


We use the keyword/phrase create table and provide a table name. The columns of
the table are defined within opening and closing parentheses and the entire statement is
ended with a semicolon. Column definitions are delimited with commas and each one
has a name, data type and a series of options that we’ll discuss later. In order to have
something to work with, we provide a full example as Code Sample 3.1.


Style

Typically, the naming rules for table names and column names are less strict than
many programming languages. Most databases allow you to use special characters,
combinations of numbers, lower and uppercase letters, underscores, and even spaces.1
Some databases treat these as case insensitive and others may treat them as case sensitive.
Or, it may even be dependent on the configuration of a particular database setup. In
 1
     To include spaces, you typically need to encapsulate table/column names using backticks:
       `My Favorite Books` . You can sometimes even used otherwise reserved SQL keywords for table
      names as long as you escape them using backticks. However, this is generally discouraged.




10
                                                                               3.2 Tables

any case, to avoid many of these issues we’ll follow a “modern” naming convention of
upper camel casing for table names and lower camel casing for column names. We prefer
this style because ultimately, our data will be associated with objects and variables in a
programming language. Following the same naming convention makes the one-to-one
connection between data and objects convenient and straightforward.
Many style guides prefer a more traditional naming convention of lower casing for table
names and lower underscore casing for column names. For example, the table in Code
Sample 3.1 using this convention might have a table name of book while the first and
last columns would have the names book_id and num_copies instead. If you have a
choice of convention, whichever you choose is fine as long as you are consistent throughout
all of your tables.
Another notable item in our styling is that we use the modern convention of using lower
case SQL keywords. Though table names and columns may or may not be case sensitive,
SQL keywords such as create table or varchar or not null are case insensitive.
The more traditional way of writing SQL is to use all uppercase letters in order to make
the SQL keywords distinct from the table/column names. For example, the SQL in Code
Sample 3.1 may instead look like the following.


  1   CREATE TABLE Book (
  2     book_id INTEGER PRIMARY KEY AUTO_INCREMENT NOT NULL,
  3     title VARCHAR(255) NOT NULL,
  4     author VARCHAR(255),
  5     isbn   VARCHAR(255) NOT NULL DEFAULT '',
  6     dewey FLOAT,
  7     num_copies INTEGER DEFAULT 0
  8   );



However, we’ll prefer the modern convention of using lower casing for all of our SQL
keywords. This has the advantage of not having to hold the shift key down while typing
half of our characters, saving our pinky fingers in the process. Moreover, the traditional
convention dates back to a time when monitors were monochromatic and using different
casing made SQL code easier to read. We live in the modern era with IDEs and syntax
highlighting (as in our examples, SQL code is highlighted in green) which arguably makes
the upper casing convention obsolete.
In addition, it is best practice to avoid pluralizations in table names. After all, it is
the Book table, not the “books table.” The table contains book records, but the table
itself is a single entity. English is not the most consistent language when it comes to
pluralization anyway. Best to avoid these issues altogether by consistently using singular
forms for all of your tables.



                                                                                        11
3 Relational Databases

Basic Data Types

Most SQL databases support a wide variety of column data types. For simplicity, we’ll
restrict our attention to a few of the most useful data types.

      • integer (or int ) is typically a traditional 32-bit signed integer allowing you
        to store values in the range [−2, 147, 483, 648, 2, 147, 483, 647]. You can insert
        and modify data within this range. However, if you try to insert a value outside
        this range, your query may result in an error or a warning. Depending on the
        particular database, your data may end up getting modified; a value greater than
        the maximum value may get rest to the maximum value for example.
         Some database systems support variations on this basic integer data type. For
         example, MySQL also supports tinyint , smallint mediumint and bigint
         which are 1, 2, 3, and 8 byte integer values respectively.
      • double or float (or real ) are floating-point numeric types. The exact sup-
        ported keywords are highly dependent on the database being used, but typically
        implement the same floating-point types as most programming languages (as IEEE
        754 floating-point numbers). double is usually an 8-byte number offering 15-17
        digits of precision and float is a 4-byte floating-point number giving 6-8 digits of
        precision.
         Many databases allow custom precision floating-point datatypes using syntax such
         as decimal(p,s) or numeric(p,s) where p is the precision (number of total
         digits of precision) and s is the scale (number of factional digits of precision or
         number of digits to the right of the decimal place). For example, decimal(65,30)
         would give you 65 total digits of precision with up to 30 digits in the factional part.
         Though it may seem like you can have arbitrarily large numbers, many databases
         will still limit the maximum precision either internally or by configuration.
      • varchar(n) can be considered as the basic string type. The keyword stands
        for variable character column in which you can store basic ASCII characters
        (or Unicode if the database supports it). The n is typically specified as the
        maximum number of characters that can be stored in the column. For example,
         varchar(255) 2 would allow the column to store any text value of up to 255
        characters. Attempts to insert or modify data that exceeds the limit usually result
        in truncation: the first n characters of the data will be stored, but any data
        exceeding the limit is cutoff and lost. Truncation may also result in a warning.
         In SQL, string literals can be denoted with either single or double quotes: 'foo'
          "bar" and 'hello world' are all examples.
 2
     You’ll often see the magic number 255 pop up in many examples as a column size that is “big enough”
      for many columns. This number has some historic meaning: it was the internal limit in earlier
      MySQL versions. The 255 would also allow the database to use at most one byte as metadata to
      store the actual length of the stored string.




12
                                                                               3.2 Tables

      Some databases include extended support for “large” text fields. MySQL for
      example supports text (64KB), mediumtext (16MB), or longtext (4GB) while
      PostgreSQL supports text which allows an arbitrarily large text field.
      There may be other issues involved with varchar . Collation refers to how data is
      stored and compared in a database. For example, varchar data may be stored
      as ASCII or Unicode or some other character set. When retrieving or ordering
      data, comparisons may be made in a case sensitive or insensitive matter. The
      configuration of the database, or individual table or column or even how a query
      is made may all determine how data is stored or compared or and may result in
      unexpected or unintended results. For this reason, it is essential to choose sensible
      defaults and to be consistent about them.
In addition to those described, databases may support many other column data types or
aliases. Aliases are simply convenient names or abbreviations for other data types. As
mentioned above, int is an alias for integer for example.
Databases may also support other non-numeric data types such as date/time values
(though these are rarely consistent and their use is often discouraged) or binary types
(often referred to as Binary Large Object (BLOB)) that allow you to store binary files
such as image, audio, and video files directly in a database table. Some databases support
enumerated types such that column values may be limited to a few predefined values or
even boolean types that may only be set to true or false . PostgreSQL even allows
you to store JSON formatted data as a column type and even allows you to query these
this sub-formatted data directly!
Databases may support each of these types simply as an alias to another simpler type.
For example MySQL supports boolean but does not have an actual boolean type.
Instead, it uses a tinyint value and uses the convention that 0 is false and any non-zero
value is true. For more details you would need to Read The Manual (RTM) of your
database system as many of these data types are non-standard and may not be portable
from one database to another.


Options

Depending on the type of data stored in a column, you may be able to apply several
options to specify additional rules or constraints to the values in the column. As with
data types, there are dozens of different options, but we’ll only focus on the most useful
here.
   • primary key auto_increment defines a primary key that is automatically as-
     signed by the database system for new records. We’ll cover keys in more detail in
     Section 3.2.2.
   • not null specifies that column values are not allowed to be null. By default,
     column values are nullable which means that specific data values do not need to be



                                                                                        13
3 Relational Databases

        specified and instead can take on a value of null (which is a recognized keyword
        in SQL). Nullable fields essentially make a column value optional and allow for
        situations in which data may be “missing” or undefined. To disallow this, you can
        use the option not null ; any attempt to insert a record without specifying the a
        valid value for the column (or alternatively, any attempt to change a column value
        to null ) will result in an SQL error and the record may be rejected. In general,
        column values should be made not null unless there is a good reason to allow
        them to be nullable.
      • default (value) allows you to define a default value for a column value. If a
        user inserts a new record without defining a value for a column with this option,
        then the default value will be used for this new record. Of course, if the user inserts
        a new record with a specified value then that value will be used instead. This
        option is typically used in conjunction with the not null option though they do
        not have to be used together.

To illustrate the effect of these options reconsider the example given in Code Sample 3.1,
repeated here for convenience. Let’s make some observations about each of the columns.

  1    create table Book (
  2      bookId integer primary key auto_increment not null,
  3      title varchar(255) not null,
  4      author varchar(255),
  5      isbn   varchar(255) not null default '',
  6      dewey float,
  7      numCopies integer default 0
  8    );

      • The title column may take on any string value of up to 255 characters, but it
        may not be set to null . Consequently any new record must have a title specified.
        Moreover, any existing record may not have its title changed to null .
      • The author column has no options so by default it is allowed to be null . New
        records inserted without an author value will default to null . Existing records
        can have their author value changed to null .
      • The isbn column is not allowed to be null but a default value is specified. In
        this case, the default is the empty string (which is not the same thing as null ),
        string literals being denoted with single or double quotes. Existing records may not
        have their ISBN value changed to null . However, new records may be inserted
        without specifying a value for the ISBN column as the database will use the defined
        default.
      • The dewey column is specified as a float (which suffices for the Dewey Decimal
        System) but with no further options, it is allowed to be null and has no default
        value. Equivalently, you can think of null as being the default value for new



14
                                                                                3.2 Tables

      records that get inserted with no dewey value specified.
   • The numCopies column is allowed to be nullable, but defines a default value
     of zero. This means that newly inserted records that do not specify a value for
      numCopies will not get set to null , but instead will get set to 0. However, you
     could modify existing records and set the numCopies column value to null if
     you wished.



Other Table Operations

Just as you can create a new table, you can also get rid of an existing table. The syntax
for doing this is simply
drop table TableName
This will remove the table entirely including all records, potentially losing a lot of data.
For this reason, many SQL clients have a “safe mode” that may be enabled by default
or may be turned on. Safe mode typically disallows you from performing potentially
dangerous operations like this. To remove a table, you may need to disable safe mode and
then drop it. This provides an extra level of safety so that you cannot absentmindedly
screw up the database.
Both create table and drop table will only work if the table does not exist/does
exist respectively. Attempts to (re)create an existing table or to drop a non-existent table
will likely result in an error. For convenience, SQL supports conditional create/drop
statements. For example,
create table if not exists TableName ...
and
drop table if exists TableName;
would only create or remove a table if they did not/did exist respectively. The queries
may still issue a warning, but not an error. The typical use case for these conditional
statements is when developing and troubleshooting a database creation script. It saves
you from having to manually re-create/re-drop tables as you debug and test your script.
Finally, you can also modify an existing table using alter table . You may want to
do this instead of dropping/recreating the table if some data exists in the table that
you do not want to lose. For details, you will need to refer to your particular database
documentation, but typically you can use syntax similar to the following.
alter table Book add column numPages int;

would add a column to the Book table as specified.
alter table Book drop column dewey;



                                                                                         15
3 Relational Databases

would remove the dewey column from the Book table as well as deleting all column
values associated with records in that table.
alter table Book modify column title varchar(100);

would change the title column in the Book table. In this case, it would shorten
the field to be at most 100 characters instead of the original 255. This would have
the consequence of truncating the column value of any record that exceeded this 100
character limit. Thus, modifying a column may also modify data stored in that column.
It also modifies the column to be nullable. If you attempt to modify the data type of a
column, the database may attempt to coerce the data types (changing numbers to strings
or vice versa) which may lead to unexpected results and data loss. For this reason, it is
always best practice to rigorously test and verify your SQL scripts and operations on
a test copy before executing them on “production data.” Even then, it is always best
practice to keep multiple (daily, weekly) database backups or snapshots.




3.2.2 Primary Keys

In our book example (Code Sample 3.1), the first we column we defined was bookId
and we designated it as a primary key . A primary key (often referred to as Primary
Key (PK)) is an identifier for records. Designating a primary key column ensures that all
records in that table have a unique value. Attempts to insert “duplicate” records with
the same primary key value will result in an error.
A primary key is important in a table because it provides a unique identity to every
record in the table and all future records as well. It is not necessary for a table to have
a primary key, but it is generally (very) poor design not to do so. Further, each table
can have at most one primary key (otherwise it would not be primary). However, tables
may have more than one key and there are several variations and other considerations
(see Section 3.2.5).
In general, a key is a way for a database to organize and retrieve records and data.
Defining a key on a column (or group of columns) means that the database will internally
organize records according to that key. This organization is known as indexing (which is
why a key is equivalently referred to as an index ). This allows a database to search and
retrieve data much more efficiently. This is similar to how binary search can be used to
speed up searching an array but only if the array is sorted (or indexed ).3 A primary key
is just a key with the additional constraint that all values are unique.


 3
     Typically databases do not simply use arrays to store data, but instead “smart” data structures such
      as B-trees.




16
                                                                                3.2 Tables

Best Practices

As previously mentioned, it is best practice to have a primary key in every table of your
database for sake of consistency, identity, and most importantly to be able to reliably
relate records between tables (see Section 3.2.3). In addition, below we identify several
other best practices when using primary keys. Though there may be a healthy debate
about the specifics, these are generally accepted as good practices.

   • Only use a single integer data column. Any data type can be designated as a
     primary key, but integer data types are best. Floating-point data types are not ideal
     because of all the problems inherent in using inexact numeric types: floating-point
     errors, round-off errors, inexact comparisons, etc. Varchar fields are also less than
     ideal for several reasons. As previously noted, varchar fields are highly dependent on
     the database, table, and query character sets used which can lead to inconsistencies
     with case sensitivity, encodings, etc. Using varchar fields can also be less efficient.
     Comparisons become string comparisons and may require comparing the entire
     field (dozens or hundreds of characters) instead of one simple number. On the
     other hand, integer types have none of these problems: they are simple, exact, and
     efficient.
   • Use auto_increment for all primary keys. Ensuring that every record has a
      unique primary key value, called key management is a very difficult problem in
      general. If you were to do this manually, you would immediately face a lot of
      difficult challenges. Before inserting any new record you’d have to check that the
     value was not already being used by any other record. Then, you’d have to “reserve”
      that value and hope that no other program, thread, or transaction came in and
      used that value before you were able to insert the new record. This is a classic
     “race condition”: two separate programs or threads may find that the key “1234” is
      unused and think it is safe to use 1234 in their new record. One of these programs
      may “win” the race and insert their record using this value. This would either end
      up in a duplicate record or an error in the second program.
     Key management is a very difficult problem that is well-solved by a relational
     database. Let it do its job and generate and assign key values for you. Using the
      auto_increment option for your primary key field ensures that the database will
     take care of the key management problem for you.
   • Do not allow primary keys to be nullable. Most databases do not allow this anyway,
     but some do. A nullable primary key is of very limited use. Since primary key
     values must be unique, at most one record can have a null primary key value anyway.
     A null primary key essentially means that the the record has no identity which is
     not only of little value but it is highly questionable if a record with no identity
     should be stored in a database to begin with. Most databases will implicitly add
      not null to a field designated as a primary key, but explicitly doing so ensures
     portable code.



                                                                                         17
3 Relational Databases

     • Use a consistent naming scheme. In following with using lower camel casing for
       fields, it is recommended that you name all of your primary key fields after the
       table name appending an Id . For example, the primary key in the Book table
       was named bookId . Following a consistent naming convention eliminates the guess
       work and the need to continually reference the database schema.
     • Use surrogate keys. Database tables often model real-world entities and often
       these entities have a natural key or a natural identity. Books, for example have an
       ISBN (International Standard Book Number) that uniquely identifies every book
       published. This would be considered a “natural key” and we may be tempted to
       use it as a primary key in our database. However, we instead used a surrogate key
       which was a number generated by our database, identifying the ISBN as a separate
       column value. Other examples may include Social Security Numbers (SSN) for
       people, NUIDs for UNL students, a SKU (Stock Keeping Unit) for products, etc.
       The reason to use surrogate keys over natural keys for your database is because
       you have little to no control over the key management of natural keys. If you don’t
       have control over the keys, then you should not allow the keys to control your
       database. Organizations can (and often do) make mistakes. It is entirely possible
       to issue the same SSN to two different people or to issue multiple student NUIDs
       to the same person. These mistakes can be rectified, but imagine having to correct
       the data in our database. If we’ve used the natural key throughout several tables
       then it may be very complicated to update the value. Natural keys may also not
       be simple integers, leading to the problems with varchar fields identified above.
       Instead, we can use auto-generated surrogate keys and avoid problems that are
       external to our database. We still have the option of storing natural keys values
       in separate columns and we can even index and make them unique (see Section
       3.2.5), but we can do so without hamstringing our database design with potential
       problems that we have no control over.


3.2.3 Foreign Keys & Relating Tables

Once again consider the Book table in Code Sample 3.1. We identified the author of a
book using a single varchar field. This can lead to several potential data anomalies,
redundancies and other problems. For example, consider the (abridged) Book table
records in Table 3.2.
All four books were clearly written by the same author, Douglas Adams. However,
because we decided to model the author with a single varchar field, it introduced the
potential for inconsistent data representation. In some records we ordered it first name-
last name, in others it was reversed. In one record we included his middle name. These
data anomalies were possible because though each book has a unique identity ( bookId ),
authors do not. Even if we had consistently used the same name representation for
Douglas Adams across all records, we would still have the problem of data redundancy:



18
                                                                                      3.2 Tables

          bookId     title                                             author
          1          Long Dark Tea-Time of the Soul                    Douglas Adams
          2          Dirk Gently’s Holistic Detective Agency           Douglas Adams
          3          The Hitchhiker’s Guide to the Galaxy              Douglas Noel Adams
          4          The Pirate Planet                                 Adams, Douglas

                          Table 3.2: Sample Book Data With Anomalies


the same string would be repeated for every book he wrote. This would be little different
from a flat file data representation.
To give author records a unique identity, we need to define another, separate table for
authors and then relate records between the two tables. To achieve this, we need to use
Foreign Key (FK). The process of separating data out into different tables is referred to
as normalization which we discuss in detail in Section ??.
A foreign key is a column whose value matches or refers to a value in a column in another
table. Generally, let A and B be tables. Further, let B have a foreign key column b
that references a column a in A. This sets up a parent-child relationship between the
two tables. We say that A is the parent table and B is the child table. Several records
in table B may have the same foreign key value b, meaning that one parent can have
multiple children. This defines a one-to-many relationship between the two tables.
To make this more concrete, let’s reconsider the relationship between book and author
data. For simplicity we’ll consider the situation in which one author may write multiple
books.4 In this scenario, there is a potential one-to-many relationship between an author
and her books. Equivalently, there is a many-to-one relationship between books and
an author. We say that an author table should be the parent table and the book table
should be the child table, so we need a foreign key in the book table that refers back to
the author that wrote it.
Let’s rewrite a simplified Book table and introduce a new Author table as follows.

     1   create table Author (
     2     authorId int primary key auto_increment not null,
     3     firstName varchar(255) not null,
     4     lastName varchar(255) not null
     5   );


     1   create table Book (
     2     bookId int primary key auto_increment not null,
     3     title varchar(255) not null,
 4
     Of course one book may have multiple authors, but we’ll save that for later.




                                                                                             19
3 Relational Databases

                                                     Author Table
                                              Douglas Adams




                                                                                       Book Table
            Long Dark Tea-Time...   Dirk Gently...      Hitchhiker’s Guide...   The Pirate Planet



Figure 3.3: A parent-child relationship between the Author and Book tables. Book
            records in the child table, refer back up to the parent Author record via a
            foreign key.


  4     authorId int not null,
  5     foreign key (authorId) references Author(authorId)
  6   );


We’ve included an authorId column in the Book table that is designated as a foreign
key that refers to a unique record in the Author table. We’re guaranteed that the
record is unique because the foreign key references the primary key. Foreign keys are
not required to refer to a primary key but it is best practice to do so (foreign keys are
required to refer to a unique field, however). Note the syntax for defining a foreign key:
we use the keywords foreign key followed by the column in the table designating the
foreign key (surrounded by parentheses). We then specify which table and which column
in that table to which the foreign key refers. A visualization of this relationship can be
found in Figure 3.3.
Once this many-to-one relationship is defined and author data and book data has been
separated, we can now see how the potential for data anomalies and redundancy is
reduced. Consider several authors and several book records as in the following two tables.
Now each author has a unique record and thus identity ( authorId ). The first name and
last name are separated and there is only one instance of each for each author solving
our representation problem. In the book table, the author is represented by the primary
key value of the corresponding author record. Arguably, there is still some redundancy
here, but it is far less. Only a single number is repeated rather than an entire string (or
other data that we may want to include relevant to an author).
Though Figure 3.3 depicts a parent-child relationship between the two tables, database
tables are usually represented using an Entity Relation (ER) diagram which has several
established conventions for visualizing tables and their relationships. An example can be
found in Figure 3.4.
Though tools that generate ER diagrams can vary in specifics, in this example each
table is represented separately. Primary keys are indicated with a key graphic, foreign



20
                                                                              3.2 Tables




                          authorId    firstName     lastName
                          42          Douglas       Adams
                          8           Terry         Pratchett
                          17          Neil          Gaiman
                          23          Cory          Doctorow
                          97          Octavia       Butler

                          Table 3.3: Normalized Author Table


            bookId    title                                      authorId
            1         Long Dark Tea-Time of the Soul             42
            2         Dirk Gently’s Holistic Detective Agency    42
            3         The Hitchhiker’s Guide to the Galaxy       42
            4         The Pirate Planet                          42
            5         Anansi Boys                                17
            6         Neverwhere                                 17
            7         Coraline                                   17
            8         Color of Magic                             8
            9         Small Gods                                 8
            10        Kindred                                    97

                          Table 3.4: Normalized Author Table




Figure 3.4: An Entity-Relation Diagram for the Author/Book tables indicating a one-to-
            many relationship. The relation is represented by a connection between the
            two tables with the vertical lines representing a “one” and the “chicken foot”
            representing “many”.




                                                                                      21
3 Relational Databases

keys are represented with red diamonds, and the data type of each column is included.
A diamond that is filled in indicates that the field is not nullable (likewise, non-filled
diamonds indicate nullable fields). Finally, the one-to-many relation is indicated with the
line between the tables with the one symbol (two vertical lines) and the many symbol
(the chicken foot) on each table respectively.
Finally, take note of the naming conventions we’ve used: the foreign key column name
exactly matches the primary key name to which it refers. Using this consistent naming
scheme will reduce the guesswork when we start formulating queries to our database.


Foreign Key Constraints

Consider again the data in Tables 3.4 and 3.3. Though we’ve defined a one-to-many
relationship, that does not necessarily imply that every record has a relationship. For
example, there is a record for the author Cory Doctorow, but he has no associated book
records. Similarly, Octavia Butler only has one associated book record. Nevertheless,
there is still a potential one-to-many relationship even if we do not have multiple records
stored in the book table.
A foreign key represents a constraint in the relation of data that must always be satisfied.
For example, suppose we were to insert a new book record for the novel Station Eleven into
the book table but without a corresponding author (Emily St. John Mandel) record in the
author table. This has several immediate problems. First, the foreign key authorId in
the book table is not nullable, so inserting such a record would fail. From the parent-child
relationship perspective, we would have a child record (book) without a parent (author),
giving us an “orphan” record. This violates the relationship that we’ve defined between
the two tables. For a child record to exist, the parent record it refers to must exist first
so that we have a valid primary key value for our foreign key to refer to. If we were to
delete records (say we wanted to remove the Douglas Adams record from the Author
table) we must do so in reverse. In order to remove a parent record, all of its children
records that refer to it must be deleted first.
This order also applies when creating tables. Since the Book table refers to the Author
table, we must create the Author table first. Likewise, if we were to drop these tables,
we must do so in reverse order: the child table must be removed before the parent table.
Attempts to manipulate data or tables in the wrong order will result in database errors.


3.2.4 Many-To-Many Relations

Let us consider extending our two table database further. It is entirely possible for one
book to have multiple authors. For example, Good Omens was written by both Terry
Pratchett and Neil Gaiman. Our current model only allows for a one-to-many relationship
between authors and books. What we really want is a many-to-many relationship. This



22
                                                                              3.2 Tables


  1   create table Author (
  2     authorId int primary key auto_increment not null,
  3     firstName varchar(255) not null,
  4     lastName varchar(255) not null
  5   );
  6

  7   create table Book (
  8     bookId int primary key auto_increment not null,
  9     title varchar(255) not null
 10   );
 11

 12   create table AuthorBook (
 13     authorBookId int primary key auto_increment not null,
 14     authorId int not null,
 15     bookId int not null,
 16     foreign key (authorId) references Author(authorId),
 17     foreign key (bookId) references Book(bookId)
 18   );

            Code Sample 3.2: A many-to-many Author/Book table design


would enable one author to be associated with several books and one book to be written
by several authors.
A one-to-many relation from table A to table B is the same thing as a many-to-one
relation from table B to table A. To enable a many-to-many relationship we define a
third table called a join table between these two tables. We then define a one-to-many
relation from table A to the join table and a one-to-many relation from table B to the
join table, resulting in a many-to-many relation between tables A and B. We do this by
defining two foreign keys in the join table. From the parent-child relation perspective,
the join table is a child of both tables (and so has two parents). This modified design is
presented as Code Sample 3.2 and the resulting ER diagram is depicted in Figure 3.5.
As with a simple one-to-many relationship, all of the same rules and constraints apply.
In this case, both an author and a book record must exist before you can associated
them with each other by inserting a record in the join table. To delete data in either
table, this association record in the join table must be removed first. We’ve named our
join table AuthorBook after the two tables that it “joins” together. Alternatively, we
could have named this table after the relation it represents, WrittenBy for example.
Data for the books from a previous example using this new many-to-many relation model
is depicted in Figure 3.6. We can now model the fact that Good Omens was written by
both Terry Pratchett and Neil Gaiman. Entries 11 and 12 in the AuthorBook join table
have the same bookId corresponding to Good Omens but different authorId values.



                                                                                       23
3 Relational Databases




             Figure 3.5: A many-to-many relationship using a join table.


The remaining entries remain similar to our old design, modeling that each of the other
book records only have one author each. Another consequence of this design change is
that we can now have book records without having to have an author record, similar to
how we could have an author with no book records before. In particular, Confederacy of
Dunces has no join record indicating its author.


3.2.5 Other Keys

Primary keys and foreign keys are two specific types of keys. You can define a more
general key on any column in any table of even a combination of columns in any table.
Equivalently, a key is referred to as an index which makes the database maintain an
ordering on the column values for fast search and retrieval.
Suppose that you’ll be making frequent search queries to the author table based on (say)
the last name. For example, you may enable users to search for authors whose last name
begins with “T” or to perform a (partial) keyword search. Placing a key or index on the
 lastName column can greatly increase performance by several orders of magnitude. To
do this, you can use the following syntax:
key (columnName)

Or, equivalently, index (columnName) ( key and index are synonymous) and the
database will keep the data (internally) organized according to values in that column.
Keys can also be defined for a combination of columns. Such keys are called composite
keys because they are composed of more than one column. To define such a key you can
simply provide a comma delimited list of column values. For example:



24
                                                                                         3.2 Tables




                                              bookId   title
   authorId   firstName   lastName
                                              1        Long Dark Tea-Time of the Soul
   42         Douglas     Adams
                                              2        Dirk Gently’s Holistic Detective Agency
   8          Terry       Pratchett
                                              3        The Hitchhiker’s Guide to the Galaxy
   17         Neil        Gaiman
                                              4        The Pirate Planet
   23         Cory        Doctorow
                                              5        Anansi Boys
   97         Octavia     Butler
                                              6        Neverwhere
                                              7        Coraline
                                              8        Color of Magic
                                              9        Small Gods
                                              10       Kindred
                                              11       Good Omens
                                              12       Confederacy of Dunces




                 authorBookId    authorId   bookId
                 1               42         1
                 2               42         2
                 3               42         3
                 4               42         4
                 5               17         5
                 6               17         6
                 7               17         7
                 8               8          8
                 9               8          9
                 10              97         10
                 11              8          11
                 12              17         11


Figure 3.6: Normalized Author/Book table data with a many-to-many relation.




                                                                                                 25
3 Relational Databases

key (columnA,columnB)
Finally, you can make keys or composite keys unique without forcing them to be primary
keys (primary keys are unique by definition). Recall that we made the design decision to
use surrogate keys in our book table instead of the “natural” ISBN key. However, we
may want to ensure that all ISBN values are unique so that we don’t have two book
entries with the same ISBN. To do this, we can use the following syntax:
unique key (columnName)
We can also combine these two concepts and put a unique composite key on a combination
of columns; for example:
unique key (columnA,columnB)
Let’s apply these features to our author-book database. The final version can be found
in Code Sample 3.3. Observe that we’ve placed a key (index) on the lastName column
for performance. We also placed a unique key on the isbn column in the Book table
to prevent duplicate entries.
We also placed a unique composite key on the combination of foreign keys in the
 AuthorBook join table. This creates a constraint so that only one book/author join
record is possible. Without this constraint we might be able to insert multiple author
records for the same book:

                          authorBookId      authorId    bookId
                          1                 42          1
                          1                 42          1
                          1                 42          1
                          1                 42          1

Four records modeling the fact that Douglas Adams wrote Long Dark Tea-Time of the
Soul are redundant and not necessary.


3.3 Structured Query Language

A relational database provide a means for storing, representing, and modeling data. We
still need a way to interact with and operate on that data. In order to interact with
data SQL provides queries for basic Create-Retrieve-Update-Destroy (CRUD). These
four operations are sufficient to perform any task we may want to on our data. Each
operation is supported by the query statement keywords insert , select , update ,
and delete respectively (though ISUD is a less compelling acronym).
To illustrate examples for each of these queries, we’ll frequently refer to the Author-Book
database as presented in Figure 3.7.



26
                                                 3.3 Structured Query Language




 1   create table Author (
 2     authorId int primary key auto_increment not null,
 3     firstName varchar(255) not null,
 4     lastName varchar(255) not null,
 5     key (lastName)
 6   );
 7

 8   create table Book (
 9     bookId int primary key auto_increment not null,
10     title varchar(255) not null,
11     isbn varchar(100) not null,
12     numCopies int not null default 0,
13     key (title),
14     unique key (isbn)
15   );
16

17   create table AuthorBook (
18     authorBookId int primary key auto_increment not null,
19     authorId int not null,
20     bookId int not null,
21     foreign key (authorId) references Author(authorId),
22     foreign key (bookId) references Book(bookId),
23     unique key (authorId,bookId)
24   );

               Code Sample 3.3: Full Author/Book table database




                                                                           27
3 Relational Databases




                       Figure 3.7: Final Author-Book Database


3.3.1 Creating Data

Before we can process data, we have to actually create some to work with. To do this,
we can insert new records into a table. The general syntax for doing this is:
insert into TableName (columnA, columnB, ...) values (valueA, valueB, ...);
We identify the table we wish to insert a record into and then provide a comma-delimited
list of column names we wish to specify values for. Another comma-delimited list of
values (string literals or numbers for example) which has a one-to-one correspondence
with the columns identified, thus the order in which we specify columns matters. These
are often referred to as tuples: a group of data delimited by commas and enclosed in
parentheses which correspond to a single record in a table.
As a concrete example, let’s insert a new record into our Author table for Isaac Asimov:
insert into Author (firstName, lastName) values ("Isaac", "Asimov");
The order of the columns doesn’t matter, but the correspondence with the values does.
Equivalently we could have written:
insert into Author (lastName, firstName) values ("Asimov", "Isaac");
and it would have resulted in the same record. Recall that neither field was nullable and
no default values were specified, so if we omitted one of them, say:
insert into Author (lastName) values ("Asimov");



28
                                                         3.3 Structured Query Language

it may result in an error or a warning and may result in either no record being inserted or
a system-wide default value assigned to the firstName field depending on the database
and configuration.
In these examples we omitted a value for the primary key authorId . As a result the
database will automatically generate and assign a unique value for this field for us. If,
instead we wished to specify a value for the primary key we could do so:

  1   insert into Author (authorId, firstName, lastName) values
  2     (1920, "Isaac", "Asimov");


Recall, however, that key management is best left to the database. Nevertheless there
are use cases for (manually) defining your own keys. For example, when you insert test
data and need to make associations (foreign keys) between tables. Providing your own
key values makes this much easier. For example, let’s also insert a record for his novel
Pebble in the Sky (omitting the ISBN for brevity):
insert into Book (title, isbn) values ("Pebble in the Sky", "0765319136");
Now we want to associate these two records with each other. To do so, we need to insert
a record into the AuthorBook table which requires both the authorId and bookId
as foreign keys. The problem is we don’t know what the bookId is as it was generated
by the database. The more verbose (and fragile) solution would be to use select
statements (see Section 3.3.2) in nested queries which might look something like the
following.

  1   insert into AuthorBook (authorId, bookId) values (
  2     (select authorId from Author where lastName = "Asimov"),
  3     (select bookId from Book where title = "Pebble in the Sky")
  4   );


This solution essentially pulls the generated key values from previously inserted records.
It is fragile because it assumes the data has successfully been created (and still exists)
and that it is unique. That is, it assumes there is one and only one author whose last
name is Asimov. If those assumptions do not hold, this query will fail.
Instead, we could provide our own key values as follows.

  1   insert into Author (authorId, firstName, lastName) values
  2     (1920, "Isaac", "Asimov");
  3

  4   insert into Book (bookId, title, isbn) values
  5     (123, "Pebble in the Sky", "0765319136");
  6

  7   insert into AuthorBook (authorId, bookId) values



                                                                                        29
3 Relational Databases


     8     (1920, 123);


This still runs the risk that the key values are not already being used by other records in
the database which is why providing your own key values should be restricted to test
data or to initialize a database with some “seed” data.
If you want to insert multiple records with a single query you can provide a comma-
delimited list of tuples:

     1   insert into Author (firstName, lastName) values
     2     ("Charles", "Dickens"),
     3     ("Jane", "Austen"),
     4     ("Frank", "Herbert"),
     5     ("Robert", "Heinlein");


In some of the above examples you’ll note that we saved on horizontal space by breaking
the query to the next line and using indentation. As with most programming languages,
whitespace does not matter and so breaking long queries up into multiple lines is good
style.
Recall that string literals can be denoted with either single quotes or double quotes.
Good style would have you choose one and use it consistently, not mixing the two ways.
Either way you choose, if you need to use a single or double quote in your string literals
you can escape them with a backslash.

     1   insert into Author (firstName, lastName) values
     2     ("Stan \"The Man\"", "Lee");
     3

     4   insert into Book (title,isbn) values
     5     ('Foundation\'s Edge', '0586058397');



3.3.2 Retrieving Data

Once a table has data in it, of course you’ll want to retrieve and process it. To retrieve
data you use a select statement which has the following syntax.
select columnA, columnB, ... from TableName;
This will retrieve the specified column values for every record in the table you are
querying.5 The collection of records this query returns is generally referred to as a
 5
     Depending on your setup, the database may use pagination so that a limited number of records are
      returned and the client will need to ask for the “next page” of results.




30
                                                         3.3 Structured Query Language

result set. Specifying a comma delimited list allows you to retrieve a subset of specific
column values that you are interested in. Alternatively, you can use the wildcard or “star”
operator to retrieve every column value:
select * from TableName;
For example, to get only the last name and first name from the Author table we could
execute the following query:
select lastName, firstName from Author;
The order of the columns has been switched in this query from how we defined the table
(though the order of columns is generally irrelevant). To retrieve all 5 columns from the
 Book table we could use:
select * from Book;
Using the wildcard operator is fine when working directly with a database via an SQL
client or when debugging/developing. Ultimately, you’ll likely want to connect to your
database programmatically. When connecting to a database programmatically, using
the wildcard is highly discouraged since it forces a lot of data to be returned in the
result set that you may not necessarily use. Since this data is typically being transmitted
over a network connection, sending unnecessary data can have a negative impact on
performance.




Aliases


When you query a table you can rename the column names in the result set by specifying
an alias. This doesn’t change the column names in the table, only in the result set. The
syntax for doing this is to use the keyword as followed by the alias. In the following
example, we query the Book database, aliasing the first two columns but not the third.


  1   select title as bookTitle,
  2          isbn as ISBN,
  3          numCopies
  4          from Book;



As we’ll see later, you can also alias table names for convenience and to simplify com-
plicated queries. In many databases the as keyword is optional. However, most style
guides have you explicitly use the keyword for column aliases while omitting it for table
aliases and we’ll also follow this convention.



                                                                                        31
3 Relational Databases

Where Clause

The the previous select queries were not limited or conditioned on anything and so
the result set contained every record in the table being queried. If you want to limit the
results of a select query, you need to qualify it with a where clause and provide some
condition on the data.
For example, suppose that we want to query all books written by Isaac Asimov (whose
 authorId is 1920). To do this we append a where clause to the end of our query with
a boolean condition.
select * from Book where authorId = 1920;
The result set of this query will only contain book records that satisfy the authorId = 1920
condition. Note that unlike most programming languages, the equality operator is only a
single quote. However, you do use the traditional != for the inequality operator. So for
example:
select * from Book where authorId != 1920;
would result in every book record not written by Isaac Asimov. Numerical and varchar
field values can use both of these operators as well as the traditional comparison operators,
 < (strictly less than), <= (less than or equal to), > (strictly greater than) and >=
(greater than or equal to). When applied to varchar fields, the comparison will use
lexicographic ordering, for example,
select * from Book where title >= "M";
will result in all books whose title starts with any letter M or later in the alphabet. Note
that lexicographic ordering is not the same thing as dictionary ordering so this query
would also include any book whose title begins with any lower case letter (since lower
case letters follow upper case letters in the ASCII text table).
The equality and inequality operators, however, cannot be used to test for nullity. Instead,
you need to use the keywords is null or is not null . For example:
select * from Book where numCopies is null;

would return all book records where the numCopies field is null.
As with other programming languages you can also combine conditions with logical
connectives using the keywords and and or as well as using parentheses and negations
to modify your conditional statements. Some examples:

  1   select * from Book where
  2     numCopies > 10 and
  3     (title <= "D" or title >= "Q");
  4

  5   -- the following are equivalent



32
                                                         3.3 Structured Query Language


  6   select * from Book where authorId != 1920;
  7   select * from Book where !(authorId = 1920);



Distinct Clause

A select query may return many duplicate entries. If we want to limit our results to only
distinct values we can use a distinct clause after the select keyword to exclude
duplicates. For example, if we queried all the last names of authors, there may be many
instances of the last name “Smith” as it is quite common. If we used the following query
 select distinct lastName from Author;
then we would only get (at most) one instance of each distinct last name in the result set.
This clause will be particularly useful when we process more complex queries in Section
3.3.5.


Like Operator

Using equality and comparison operators for varchar columns only enables us to formulate
queries for ranges of values. Often we want to make partial string comparisons. For
example, retrieving all strings that begin with a certain letter or strings that contain a
certain substring.
To perform partial string matching searches you can use the like clause in conjunction
with the string wildcard symbol, % (which is not the same as the column wildcard or
“star” operator). For example, to search for all authors whose last name begins with “A”
we would write
 select * from Author where lastName like "A%";
This string wildcard matches any string including the empty string, so this query would
also match any author record whose last name was only a single “A” character. You
can also define groups of characters similar to regular expressions. The following query
matches all authors whose last name begins with an “A” a “B” or a “C”:
 select * from Author where lastName like "[A-C]%";
or:
 select * from Author where lastName like "[ALZ]%";
would match any last name beginning with an “A” an “L” or a “Z”. You can invert the
group using an exclamation point:
 select * from Author where lastName like "[!ALZ]%";
would match all records that don’t begin with an A, L or Z. This is necessary as there is



                                                                                        33
3 Relational Databases

no not like clause.
You can place the wildcard anywhere you like so that
select * from Author where lastName like "%e";
matches last names that end in “e” and
select * from Author where lastName like "%e%";
match last names that contain an “e”. You can expand the string value so that you can
perform substring searches.
select * from Author where lastName like "%the%";
will match any record whose last name contains the entire character sequence “the”.
The string wildcard, % matches any number of characters (including none). To match
exactly one character you can use the underscore wildcard which matches any single
character.
select * from Author where lastName like "_s%";
will match all last names that begin with any letter but that are immediately followed by
a lowercase “s” (followed by any other string). You can use any number of combinations
of character wildcards, string wildcards, and string literals to do partial string matching.


In Operator

If you have a lot of values that you want to match you could write a very large query
with a lot of or conditions:

  1   select * from Author where
  2     authorId = 1920 or
  3     authorId = 3213 or
  4     authorId = 5345 or
  5     authorId = 4324;


This can be written in a more concise manner using the in operator and specifying a
comma delimited list of values:

  1   select * from Author where
  2     authorId in (1920, 3213, 5345, 4324);


which works equally well for varchar fields.




34
                                                         3.3 Structured Query Language


  1   select * from Author where
  2     lastName in ("Asimov", "Clarke", "Ellison");


The in operator is especially useful when you combine it with nested queries. For
example:
  1   select * from Author where
  2     authorId in (select authorId from AuthorBook);


would give us all authors who have at least one corresponding record in the AuthorBook
table. As we’ll see later, this is better done with a join query, but you can also negate
the operator to get all author records that do not have any corresponding book records:
  1   select * from Author where
  2     authorId not in (select authorId from AuthorBook);


Order By Clause

In general the order of records in a result set has no meaning. The database stores and
returns results in whatever manner is the most efficient. If you want records to be return
in a specific order you can do so by appending an order by clause and specifying which
column or columns to order the data by in either ascending ( asc ) or descending order
( desc ).
Some examples:

  1   -- order authors by last name in ascending order (default)
  2   select * from Author order by lastName;
  3

  4   -- order authors by last name in ascending order (explicitly)
  5   select * from Author order by lastName asc;
  6

  7   -- order authors by last name in descending order (Z to A)
  8   select * from Author order by lastName desc;
  9

 10   -- order authors by last name then by first name if they
 11   -- have the same last name
 12   select * from Author order by lastName, firstName;
 13

 14   -- order books by most number of copies first (descending),
 15   -- then by title (ascending)
 16   select * from Book order by numCopies desc, title;




                                                                                       35
3 Relational Databases

3.3.3 Updating Data

Records that already exist in a database can be updated using an (surprise!) update
statement. The general syntax is as follows.
update TableName set columnA = valueA, columnB = valueB, ... where [expression]

The where clause part of the query is extremely important as it will limit the effect of
our statement.
As an example, suppose that we messed up the insertion of Isaac Asimov by mixing up
his first/last name:

  1   insert into Author (authorId, lastName, firstName) values
  2     (1920, "Isaac", "Asimov");


To update the record we would so something like the following:
  1   update Author
  2     set firstName = "Isaac", lastName = "Asimov"
  3     where authorId = 1920;


Of course we could have omitted the where clause and executed a statement like:
update Author set lastName = "Asimov";
which would have some extreme unintended consequences. If this query were to success-
fully executed, every author record’s last name would now be “Asimov”. Most (sane)
SQL clients won’t let you do this by default. Recall that we mentioned safe mode in
which certain dangerous queries, queries that alter data without restrictions are not
allowed. If you really do intend to alter every record, you can shut off safe mode and
proceed at your own peril. The where clause in the correct example limits the effects of
the update statement to only one record. In particular, the record whose authorId is
1920. In general, any subset of column values can be changed with an update query
and the order does not matter.


3.3.4 Destroying Data

Removing data from a table can be done with a delete query. The general syntax is as
follows.
delete from TableName where [expression]

As with the update statement, the where clause limits the effect of a delete statement.
Without a where clause, it would effectively remove every record in the table. However,



36
                                                        3.3 Structured Query Language

it would not delete the table (that is done using drop table , see Section 3.2.1).

If we wanted to delete Isaac Asimov from our Author table we would use the following.
delete from Author where authorId = 1920;
Of course if there were any references to this record in the AuthorBook table, those
would need to be deleted first as the foreign key constraints would prevent us from
deleting a parent record that still had child records.


3.3.5 Processing Data

SQL not only allows us to perform basic CRUD, but it has many powerful queries that
allow you to process and aggregate data directly in the database. In this section we cover
only a small selection of these capabilities.


Aggregates

Aggregate functions are functions that allow you to summarize data such as summing
values, taking the average, finding the minimum/maximum values etc.
The count function can be used to determine how many records would be in the result
set of your query. For example,
select count(*) from Book;
would return a count of the number of book records. The result of this query is only a
single record: the number of book records. It does not include the actual book records
in the result. Further, the “column” is not an actual column in the database. In this
case the column “name” is count(*) . We can use an alias to give this column a more
sensible name:
select count(*) as numberOfBooks from Book;
You can also qualify aggregate functions with a where clause.
select count(*) as numberMissingBooks from Book where numCopies = 0;
The sum function can be used to sum values in a particular column. Some examples:

 1   -- find the total number of copies of all books
 2   select sum(numCopies) as totalNumberOfBooks from Book;
 3

 4   -- find the total number of copies of all books
 5   select sum(numCopies) as numberOfCommonBooks
 6     from Book
 7     where numCopies >= 10;



                                                                                       37
3 Relational Databases




       Figure 3.8: A 3-D cube projected onto a 2-D surface producing a square.


You can also find the average using avg :

select avg(numCopies) as averageNumCopies from Book;

Finally, you can use min and max to find the minimum and maximum values respectively.

  1   select min(numCopies) from Book;
  2   select min(numCopies) from Book where numCopies > 0;
  3

  4   select max(numCopies) from Book;


Both of these aggregate functions work for varchar fields as well:
select max(title) as lastBookTitle from Book;
would return the lexicographically maximum (last in alphabetic order) title of a book.


Projections

In mathematics you can take an n-dimensional object and remove one of the dimensions
to project that object into an (n − 1) dimensional space. For example, you can take a
3-D cube and project it down onto a 2-D plane. This is essentially removing one of the
coordinates, mapping points (x, y, z) → (x, y). Depending on the orientation of the cube
you may get a variety of shapes on the 2-D surface. One possibility results in a square as
depicted in Figure 3.8.
Recall that a data record (a single row) is called a tuple which is essentially the same thing
as a point (in fact, we even use the same notation, (x, y, z) or (author, title, numCopies) ).
With respect to data, we can perform a similar projection, removing one of the “di-
mensions” by removing one (or more) of the columns. In addition, rather than simply



38
                                                       3.3 Structured Query Language

“loosing” information as with the cube-to-square example, we can use this project to
produce more aggregate information. The way we do this is by grouping collections of
data and collapsing them into a single record. In SQL we use the group by clause.
Before we get into the specific syntax, let’s use a motivating example to understand how
it works. Suppose we have a (simplified) book table with only three columns: the book
title, author name, and number of copies of the book in our library.
The table may look something like this.

        title                                     author            numCopies
        Naked and the Dead                        Norman Mailer     10
        Dirk Gently’s Holistic Detective Agency   Douglas Adams     4
        Barbary Shore                             Norman Mailer     3
        The Hitchhiker’s Guide to the Galaxy      Douglas Adams     2
        The Long Dark Tea-Time of the Soul        Douglas Adams     1
        Kindred                                   Octavia Butler    5

Our goal is to produce a report of how many total copies of books by each author we
have in our collection. We have 3 books by Douglas Adams with 4 + 2 + 1 = 7 copies. As
we saw in the previous section we can easily get a total number of copies using the sum
aggregate function, but that would only result in one number. In this example, we want
a report of three numbers: the total for each author. Conceptually, we need to group the
records in this table by the author. Giving us the following result.

        title                                     author            numCopies
        Naked and the Dead                        Norman Mailer     10
        Barbary Shore                             Norman Mailer     3
        Dirk Gently’s Holistic Detective Agency   Douglas Adams     4
        The Hitchhiker’s Guide to the Galaxy      Douglas Adams     2
        The Long Dark Tea-Time of the Soul        Douglas Adams     1
        Kindred                                   Octavia Butler    5

After grouping the data we can then project it down by eliminating one of the irrelevant
columns (the book titles) and summing up the numCopies values for each group. That
results in the following:

                          author         totalNumCopies
                          Norman Mailer        13
                          Douglas Adams         7
                          Octavia Butler        5



                                                                                     39
3 Relational Databases

To perform this data projection in SQL, you can use the group by clause. In the
example above we could have written the following.

  1   select author, sum(numCopies) as totalNumCopies
  2     from Book
  3     group by author;


The group by clause grouped our data for us, and the sum aggregate was responsible
for projecting the data and producing the new value. Without the sum aggregate we
would only get the collapsed data records (likely with the first value in the group for
 numCopies instead of the projected sum).

Because the totalNumCopies is not a value in the original table, we cannot use a where
clause if we wanted to further winnow this data. Instead, SQL provides another keyword
to restrict the results after a data projection using a having clause. For example, we
could further restrict the result in the example above as follows.

  1   select author, sum(numCopies) as totalNumCopies
  2     from Book
  3     group by author
  4     having totalNumCopies > 5;


This would end up excluding the Octavia Butler record:


                           author        totalNumCopies
                           Norman Mailer       13
                           Douglas Adams        7


Joins

In our previous example we used a simplified table to illustrate data projection. However,
in our author-book database, the data we’re interested in is normalized and separated out
into 3 different tables. In order to bring data together from 3 separate tables, we need to
join them together (which is why we called the table between them a “join” table). To
do this, we use a join clause.
Conceptually, a join corresponds to a cartesian product of tuples in mathematics combined
with a condition by which tuples are joined. Without a condition, this corresponds to a
regular cartesian product:


                             A × B = {(a, b)|a ∈ A, b ∈ B}



40
                                                                    3.3 Structured Query Language

For simplicity, suppose that A = {10, 20} and B = {5, 15, 25} are simple sets of single
values (instead of rows). The cartesian product would consist of the following pairs:


                          (10, 5), (10, 15), (10, 25), (20, 5), (20, 15), (20, 25)


Each element in the first set is paired with each element in the second set. In general, if
the sets are of cardinality (size) n and m respectively, then there would be n · m total
elements in the cartesian product (above, we have 2 · 3 = 6 pairs in the result).
We can limit a cartesian product by adding a condition by which a pair may be included.
This corresponds to a traditional relation on sets in mathematics:


                            R ⊆ A × B = {(a, b) | a ∈ A, b ∈ B, P (a, b)}


where P is some predicate that is either true or false depending on the elements a, b. In
fact, this is why we call them relational databases. From the example above, suppose we
put a condition that the pair (a, b) be included only if a < b. This would result in the
smaller result


                                        (10, 15), (10, 25), (20, 25)


excluding the other 3 elements.
With respect to a database, a join clause combines records (rows) in one table (table
A) with records in another table (table B) usually with some condition that limits the
combinations and defines how rows in each table get paired up. Further, we can and will
perform joins with multiple tables just as you can with cartesian products and multiple
sets (for example: A × B × C would give the obvious6 result).
The result of a join is still a “table” in that it has columns and rows as the result set,
but it is not a table that exists in the database. In general you can join tables together
using any criteria you want, but by design they are usually joined together using foreign
keys. There are also several types of joins (left/right, inner/outer, cross joins, full outer
joins, etc.) but to keep things simple we’ll focus on two of the most useful: inner joins
and left outer joins.
To illustrate how this works, let’s consider the following (simplified) Author , Book ,
and AuthorBook tables with the following data. The SQL used to generate this data is
present in Appendix 1 if you want to follow along with your own database.

 6
     If not obvious, then A × B × C = {(a, b, c) | a ∈ A, b ∈ B, c ∈ C}




                                                                                              41
3 Relational Databases

                           authorId      firstName   lastName
                           1             Norman      Mailer
                           2             Douglas     Adams
                           3             Octavia     Butler
                           4             Cory        Doctorow


             bookId   title                                     numCopiess
             1        Naked and the Dead                        10
             2        Dirk Gently’s Holistic Detective Agency   4
             3        The Hitchhiker’s Guide to the Galaxy      2
             4        The Long Dark Tea-Time of the Soul        1
             5        Barbary Shore                             3
             6        Kindred                                   5


                           authorBookId     authorId   bookId
                           1                1          1
                           2                2          2
                           3                2          3
                           4                2          4
                           5                1          5
                           6                3          6


An inner join combines records from table A with records from table B based on some
criteria. This is one of the most common joins and so the inner keyword is optional
and we’ll use the abbreviated join . For example, let’s join records in our Author table
with the join table AuthorBook . By our design, records in these two tables are related
by the foreign key authorId . We could do this using the following query.

  1    select * from Author a
  2      join AuthorBook ab
  3      on a.authorId = ab.authorId;


A few things of note about this query.
      • We join from the Author table to the AuthorBook table
      • For both tables, we provide an alias (without using the as keyword) so that we
        can refer to them in the on clause



42
                                                         3.3 Structured Query Language

      • The on keyword is used to provide the condition by which records should be joined.
        Without this, each record in the first table would be paired with every record in
        the second table.
The result of this query would be a table that would look like the following.

         authorId    firstName    lastName    authorBookId     authorId    bookId
         1           Norman       Mailer      1                1           1
         2           Douglas      Adams       2                2           2
         2           Douglas      Adams       3                2           3
         2           Douglas      Adams       4                2           4
         1           Norman       Mailer      5                1           5
         3           Octavia      Butler      6                3           6

The first three columns correspond to records in the Author table and each record is
paired with its referencing record in the AuthorBook table (the last three columns). If
we were generating a report there is some redundant information that we don’t necessarily
need (the authorId is repeated for example and the bookAuthorId is not useful for
the report). We can use this as an opportunity to limit our query to only the relevant
columns.

  1    select a.firstName, a.lastName, ab.bookId from Author a
  2      join AuthorBook ab
  3      on a.authorId = ab.authorId;


In the above query we used the table aliases when specifying columns. Though it is not
necessary in this particular example, it would be necessary to disambiguate the columns
if we had repeated column names. The result would be the following simpler report:

                              firstName    lastName   bookId
                              Norman       Mailer     1
                              Douglas      Adams      2
                              Douglas      Adams      3
                              Douglas      Adams      4
                              Norman       Mailer     5
                              Octavia      Butler     6

This only gives us the internal bookId primary key value which is of little help in a
report intended for human consumption. To get the book title into this report, we need
to join to yet another table, the Book table.



                                                                                       43
3 Relational Databases

           firstName     lastName    title
           Norman        Mailer      Naked and the Dead
           Douglas       Adams       Dirk Gently’s Holistic Detective Agency
           Douglas       Adams       The Hitchhiker’s Guide to the Galaxy
           Douglas       Adams       The Long Dark Tea-Time of the Soul
           Norman        Mailer      Barbary Shore
           Octavia       Butler      Kindred


            Table 3.5: Final results for the join of our author-book database.


  1   select a.firstName, a.lastName, b.title from Author a
  2     join AuthorBook ab on a.authorId = ab.authorId
  3     join Book b on ab.bookId = b.bookId;


which gives us the final result:


Other Joins

 In the example we joined tables according to their design by using foreign key values. If
we did not use any conditions (no on clause) then we would have gotten a complete
“cross join”. For example:

  1   select * from Author join AuthorBook;


would have produced a result similar to the following.
Which is essentially a full cartesian product. Since there were 4 records in the Author
table (table A) and 6 records in the AuthorBook table (table B), there are 4 · 6 = 24
records in the “cross join” of the two tables. However, the interpretation of the data is
meaningless, not every author authored every book. This illustrates the importance of
both good database design and sensible queries.
Another mistake we could make is to join the two tables along mismatched column
values. For example, in our design there is no direct relationship between the authorId
(author primary key) and the authorBookId (the join table primary key) which are
both independently generated by the database and have no real connection with each
other. Nevertheless, if we tried to join the two tables together using these column values
as in the following query,




44
                                                         3.3 Structured Query Language

           authorId   firstName   lastName   authorBookId   authorId   bookId
           1          Norman      Mailer     1              1          1
           1          Norman      Mailer     2              2          2
           1          Norman      Mailer     3              2          3
           1          Norman      Mailer     4              2          4
           1          Norman      Mailer     5              1          5
           1          Norman      Mailer     6              3          6
           2          Douglas     Adams      1              1          1
           2          Douglas     Adams      2              2          2
           2          Douglas     Adams      3              2          3
           2          Douglas     Adams      4              2          4
           2          Douglas     Adams      5              1          5
           2          Douglas     Adams      6              3          6
           3          Octavia     Butler     1              1          1
           3          Octavia     Butler     2              2          2
           3          Octavia     Butler     3              2          3
           3          Octavia     Butler     4              2          4
           3          Octavia     Butler     5              1          5
           3          Octavia     Butler     6              3          6
           4          Cory        Doctorow   1              1          1
           4          Cory        Doctorow   2              2          2
           4          Cory        Doctorow   3              2          3
           4          Cory        Doctorow   4              2          4
           4          Cory        Doctorow   5              1          5
           4          Cory        Doctorow   6              3          6




 1   select * from Author a
 2     join AuthorBook ab on a.authorId = ab.authorBookId;


then we would get something like the following result:

       authorId    firstName      lastName   authorBookId     authorId    bookId
       1           Norman         Mailer     1                1           1
       2           Douglas        Adams      2                2           2
       3           Octavia        Butler     3                2           3
       4           Cory           Doctorow   4                2           4


If you examine the values in the Book table, the four results have nothing to do with
the results. The first two records match (coincidence), but the third and fourth record
do not: Octavia Butler did not write Hitchhiker’s Guide... and there are no book records



                                                                                     45
3 Relational Databases

at all for Cory Doctorow. Again, it is necessary to write queries that conform to the
intended design of your database.


Left Joins

The second type of join that we’ll cover is a left outer join or simply just left join .
Take another closer look at the “final” result of a regular join clause on our database
as presented in Table 3.5. A close examination will show that Cory Doctorow has a
record in the Author table but no corresponding book records. As such, his record did
not appear in the results.
With a regular join clause, records in table A that do not have a matching record in
table B (according to the criteria that you specify in the on clause) are not included in
the result set. A left join , however, will preserve records in table A even if they do
not have a matching record in table B. For example, the following modified left join
query:
  1   select a.firstName, a.lastName, b.title from Author a
  2     left join AuthorBook ab on a.authorId = ab.authorId
  3     left join Book b on ab.bookId = b.bookId;


The left join clauses ultimately preserve the Cory Doctorow record:

             firstName   lastName   title
             Norman      Mailer     Naked and the Dead
             Douglas     Adams      Dirk Gently’s Holistic Detective Agency
             Douglas     Adams      The Hitchhiker’s Guide to the Galaxy
             Douglas     Adams      The Long Dark Tea-Time of the Soul
             Norman      Mailer     Barbary Shore
             Octavia     Butler     Kindred
             Cory        Doctorow   null

We had to use a left join on both tables as there were no records in the join table
for Cory Doctorow and thus, by design, were no records in the Book table either. For
records that have no matches, the resulting column values are null .
The term left is used in this type of join because we are joining from table A to table
B. You can also use a right join if you wanted to join from table B to table A instead,
but that would be equivalent to a left join with the order of the tables reversed in
the query. It is much easier to read from left to right and so it is common to simply
reverse the tables if you want a right join and use left join exclusively.



46
                                                         3.4 Normalization & Database Design

One instance in which you may want to do both a left join and a right join at
the same time is if you want to join two tables but preserve unmatched records in both.
For example, if we wanted to preserve all authors with no book records in our join table
but also include all books with no author record in our join table. This can be done
using a full outer join query. For example:

     1   select * from TableA a
     2     full outer join TableB b on a.colId = b.colId;


In our particular design using a join table, a better approach would be to “simulate” a
full outer join using a union or union all operator.7
Consider the following query.

     1   select * from Author a
     2     left join AuthorBook ab on a.authorId = ab.authorId
     3     left join Book b on b.bookId = ab.bookId
     4   union
     5   select * from Author a
     6     right join AuthorBook ab on a.authorId = ab.authorId
     7     right join Book b on b.bookId = ab.bookId;


The first select statement is similar to our previous example. The second select
statement uses the same order but a right join to preserve book records. We use the
same ordering in both statements so that the columns will match up. Between them
we use a union operator to combine the two result sets. Though we use the keyword
 union this will not include duplicates (which is what we usually want). To preserve
duplicates from both result sets you can use union all .


3.4 Normalization & Database Design

Recall that the purpose of creating a relational database is to separate data out into
different tables to minimize the potential for data anomalies and enforce data integrity.
This process is formalized as database normalization. Informally, normalization is simply
just common sense. When you design a database you generally identify the entities that
you wish to model and design a table for each of them. Entities usually correspond to
real-world entities and have a natural “decomposition” that lend themselves to individual
columns.
In our running example we naturally identified a table for books and one for authors.
 7
     In some databases, in particular MySQL, full outer joins are not even supported, so you must use
      these operators.




                                                                                                  47
3 Relational Databases

Though these entities are related, they don’t have an is-a relationship (a book is not
an author and an author is not a book). Therefore, we designed a table for each of
them. We naturally identified the components of each (a person has a first name, last
name; a book has a title, ISBN, etc.). We then identified the relationship between these
tables that we wanted to model. From this perspective the process of normalization is
straightforward and natural.
More formally, however, there are several levels of normal form that formalize how tables
are designed to best minimize redundancy and failure points. Though several “advanced”
normal forms have been developed, we’ll focus on the original three developed by Edgar
Codd. Each normal form builds on the previous and identifies a characteristic about the
column attributes/fields in a table and their relationship to the table’s primary key.


First Normal Form

For a database to conform to first normal form, it may not have any column that
represents more than one value. Each attribute in each table must have only atomic
values that pertain to the key. Tables that violate first normal form would have column(s)
that capture a list or series of values. For example, suppose we wanted to store multiple
email addresses for a customer. Suppose further that we defined a column named emails
and stored a comma-delimited list of values in a single varchar field so as to support as
many email addresses we wanted. Such a column does not represent an atomic value,
but instead should be modeled using a one-to-many relationship between a customer and
an email. To conform to first normal form, we should separate this out into a second
table and define an appropriate foreign key.
Using a comma-delimited “list” of values causes many practical problems as well. First,
it requires additional processing to deserialize (split out individual email records) and
serialize (formatting a list with delimiters, potentially escaping delimiters, etc.) every
time you interact with records. Using a simple varchar field also precludes enforcing
data integrity (foreign keys needing to refer to an existing record) and loses semantics
(it represents elements only as strings instead of entities/tables). Cases where a record
has no referencing record or only one referencing record (no emails or only one email)
become corner cases that need to be dealt with separately (no delimited list or usage of
a flag or special value may be necessary).
In general, violating first normal form defeats one of the main purposes of using a
relational database: it removes formal relations!


Second Normal Form

For a database to conform to second normal (2NF) it must conform to first normal form
and no non-prime attribute may be dependent on any proper subset of prime attributes.
This mostly pertains to tables that use a composite key, a primary key that consists



48
                                                              3.4 Normalization & Database Design

of multiple columns. If you take the design approach of using surrogate keys, you are
essentially guaranteed second normal form.
Consider the following example that violates second normal form: a purchase record may
have a composite key consisting of both a customerId and storeId . If the table also
contained the storeAddress , this would be a violation because the store address would
only depend on the store (and thus the storeId ) and not the customer. The solution
is again to split out these two separate entities into their own tables and relate them.


Third Normal Form

For a database to conform to third normal (3NF) it must conform to second normal form
(and thus, first normal form as well) and no non-prime column is transitively dependent
on the key. Put another way, no non-prime column may depend on any other non-prime
column. All columns must only depend on the key.
An example of a table that would violate third normal form would be a purchase table.
Suppose we stored a price-per-unit, quantity, and total columns in this table. The total
value would be dependent on the other two since total = price-per-unit × quantity.
Given the other two columns we could easily derive the total without having to store it
explicitly. Storing the total can lead to data anomalies if either of the other two columns
were ever changed.
A common saying that summarizes these normal forms is that “every non-key attribute
must provide a fact about the key (1NF), the whole key (2NF), and nothing but the key
(3NF), so help me Codd.”


Denormalization

Normalization provides good general guidance when designing a database, but normaliza-
tion alone does not guarantee a perfect design nor does it guarantee good performance
in a database. Adhering to the normal forms (at least as a starting point) can provide a
good means to design tables and identify the relationships you want to model, but there
are instances in which you may want to denormalize a database for better performance
or to simplify a design.
Reconsider the email example where we want to support multiple emails. It is reasonable
that a customer or person would have more than one email address. However, it is likely
not the case that they would have (say) dozens or hundreds of emails and even if they
did, the business use case for supporting such a model may not be justifiable. As a
simplification we could hardcode a fixed number of email columns ( Email1 , Email2 ,
etc.) effectively creating a one-to-k relationship for a fixed constant k. This is a (slight)
violation of first normal form, but it eliminates the need to join to another table.8
 8
     Indeed, if you ever find yourself creating a table-to-table relationship that is essentially a one-to-one




                                                                                                           49
3 Relational Databases

Violating normal forms is not a critical design flaw, but should be done only after careful
consideration and with very good justification.


3.4.1 Design Example

Let’s apply what we’ve covered so far and design a simple database to support a course
enrollment system similar to the data we started with in Section 3.1. Our goal will be to
design a database to support a course roster system. The design should be able to model
students, courses, and their relation to each other. The system will also need to email
students about updates in enrollment.
In general, we’ll want a table for each entity in this problem. For each one we’ll want to
identify the fields (columns) that define that entity and identify the relationships between
tables. For this problem we’ll need tables for both students and courses. We’ll start with
the student table.
We’ll use the stylistic elements we identified before: UpperCamelCasing and singular
forms for tables, lowerCamelCasing for column names, and tableNameId for primary
key columns. Here’s the full SQL for this table.

  1     create table Student (
  2       studentId int not null auto_increment primary key,
  3       firstName varchar(255) not null,
  4       middleName varchar(255),
  5       lastName varchar(255) not null,
  6       nuid varchar(8) not null,
  7       --ensure that NUIDs are unique:
  8       constraint `uniqueNuids` unique index(nuid)
  9     );


We’ve made both the firstName and lastName required by making them non nullable.
However, the middleName is nullable, making it optional (some cultures do not have
middle names). We’ve also included an nuid (Nebraska University ID) as natural key
but have not made it our primary key. We’ve done this by giving it an index and a
uniqueness constraint. We’ve created a surrogate key (auto generated studentId field)
as our primary key instead.
To support multiple email addresses for students, we’ll model a full one-to-many relation-
ship, which requires (by 1NF) a separate email table.

  1     create table Email (
  2       emailId int not null auto_increment primary key,

      relationship you may want to rethink your design and collapse the columns to one table.




50
                                                  3.4 Normalization & Database Design


  3     studentId int not null,
  4     address varchar(255) not null,
  5     foreign key(studentId) references Student(studentId)
  6   );


We’ve used the convention that the foreign key has the same name as the key it references.
Furthermore, its type (int) must match the key type it references (yet another reason to
prefer integers for primary keys). Keep in mind that order is important here: since the
Email table references the Student table, we have to create the Student first.
A course table is straightforward and we’ve included larger text field to support full
course descriptions.

  1   create table Course (
  2     courseId int not null auto_increment primary key,
  3     subject varchar(4) not null,
  4     number varchar(4) not null,
  5     title varchar(255) not null,
  6     description varchar(4096)
  7   );


We still need to model the actual enrollment of students into courses. Clearly this is a
many-to-many relationship and we should thus create a join table to bring them together.
We’ll use two foreign keys referencing a student and a course. However, we also want to
model when the student enrolled in the course. For this we can include a semester column;
but what type should it be? We could make it a varchar to support representations
like “Fall 2019” or “Spring 2020”. However, such lexicographic representations are not
well-ordered. When sorted alphanumerically, “Fall 2019” would come before “Spring
2020” even though they are out of order temporally. It is common to instead establish an
encoding that uses integers (which are well-ordered). These are often referred to as sort
keys. The drawback to this is that the encoding is not necessarily readable to end-users
and should/must be converted appropriately.

  1   create table Enrollment (
  2     enrollmentId int not null auto_increment primary key,
  3     studentId int not null,
  4     courseId int not null,
  5     semester int not null,
  6     foreign key(studentId) references Student(studentId),
  7     foreign key(courseId) references Course(courseId)
  8   );


There is still a problem though. There is nothing that prevents us from inserting multiple



                                                                                       51
3 Relational Databases




                            Figure 3.9: Enrollment Database


records for the same student, same course, in the same semester. Though a student could
take the same course multiple times (say if they withdrew the first time), they can’t take
the same course in the same semester. This should be considered bad data and we should
define a constraint to prevent it. Specifically we could add a uniqueness constraint on
the combination of all three of these fields:
constraint `uniqueEnrollment` unique index(studentId,courseId,semester)
There are lots of other things we could add and model (offerings, instructors, etc.) and
many more rules and constraints we could define. We leave these as Exercise 3.4. The
final ER digram for this database is presented as Figure 3.9.
The same data as in Table 3.1 is depicted as an ER Digram in Figure 3.10. Observe that
we’ve now solved most of the problems we identified that motivated using a relational
database. Each piece of data is now organized and have specific types. There is minimal
duplication of data. Entities are represented by unique IDs, ensuring identity (Tom
Waits is now the same as t. Waits). Data integrity rules are now enforced (students only
have one unique NUID) and relationships are well-defined.



3.5 Miscellaneous

Data, data representations, and databases are a huge topic that could not only span a
course but an entire series of courses. We’ve only touched on the core topics there are
dozens of other issues with data and databases that are beyond the scope of this chapter.
In actual organizations, there may be entire teams devoted to Database Administrator
(DBA). DBAs handle data maintenance, backups, migration, redundancy and reliance,
quality assurance, tune performance, etc.



52
                                                                                                                                   3.6 Exercises

                               studentId   firstName   middleName   lastName   nuid               courseId     subject   number   title
                               1           Tom         Alan         Waits      10001949           42123        CSCE      156      Computer Science II
                               2           Lou         Allan        Reed       10001942           12333        CSCE      230      Computer Hardware
                               3           John        null         Student    12345678           11132        CSCE      235      Discrete Mathematics
                               4           Philip      J.           Fry        30001974           32132        MATH      479      Wonton Burrito Meals



   emailId   studentId   address
   8         2           reed@gmail.com
   10        1           tomwaits@hotmail.com
   13        1           twaits@email.com
   14        3           jstudent@unl.edu                  enrollmentId   studentId   courseId   semester
   16        4           fry@unl.edu                       1              1           42123      Fall 2019
   17        3           jstudent@cse.unl.edu              2              1           11132      Spring 2020
                                                           3              3           42123      Fall 2019
                                                           4              3           11132      Fall 2019
                                                           5              4           32132      Fall 2017
                                                           6              2           42123      Fall 2018
                                                           7              3           12333      Fall 2019
                                                           8              2           11132      Spring 2020



Figure 3.10: Enrollment Data (some fields have been removed for presentation and others
             have been modified to be more readable)


With respect to relational databases, there are dozens of other commands, operations
and features that we’ve not covered here such as temporary tables, views, triggers,
stored procedures and more. There is also the issue of programmatically connecting to
a database which most programming languages support either natively or through a
library. Java for example provides the Java Database Connectivity API (JDBC) which
defines interfaces to interact with a generic database. You write Java code to connect to
a database, execute queries (possibly) receive a result set back to process. Using and API
and interfaces like this is a form of dependency inversion as discussed in Section ??. Many
programming languages also provide Object-Relational Mapping (ORM) functionality in
which you can define a mapping from your objects/member variables to tables/columns
in a database and let the library generate the queries for you. Java’s Java Persistence
API (JPA) is one such example.
Beyond relational databases are non-relational databases, often called No-SQL that
offer performance improvements at the cost of the ACID principles. Databases such as
Apache’s Cassandra and Amazon’s DynamoDB store data as Key Value Pair (KVP)
instead of related tables. Redis is a popular general purpose database that allows you to
store data as data structures (so entire list or Tree data structures are directly stored in
the database and can be queried).



3.6 Exercises

Exercise 3.1. Consider a database to model a video game library as depicted in the ER
diagram in Figure 3.11. As designed, video games have a single publisher (a publisher
may have published more than one game) and games may be available on more than one
platform (while platforms may certainly have more than one game).



                                                                                                                                                         53
3 Relational Databases




                         Figure 3.11: Simple Video Game Database


Reverse engineer this database and write SQL code to create all four tables. Use a tool
to reproduce the ER diagram and identify and fix any mistakes.

Exercise 3.2. Write SQL queries to produce the following reports for data from the
video game database.
     1. List all video games in the database
     2. List all video games that start with “G”
     3. List all publishers in the database
     4. List all video games along with their publishers
     5. List all video games along with their publishers, but only the relevant fields
     6. List all publishers in the database along with all their games, even if they don’t
        have any
     7. List all publishers with a count of how many games they have
     8. List all games and all systems that they are available on
     9. List all games that are not available on any system
 10. List the oldest game(s) and its platform(s)
 11. Flatten the entire data model by returning all data on all games
 12. Insert a new game, Assassin’s Creed, published by Ubisoft
 13. Make the new game available on at least two platforms
 14. Update the record for Megaman 3: the publisher should be Capcom, not Eidos
 15. Delete the publisher Eidos



54
                                                                         3.6 Exercises

 16. Write a query to return all games along with the number of platforms they are
     available on

Exercise 3.3. Add constraints to your video game database to prevent multiple game
records for the same platform (in the same year).

Exercise 3.4. Modify the enrollment database so that instead of a single course table,
there are specific offerings for each semester that students can enroll in. Add support
for instructors that teach particular offerings of courses.

Exercise 3.5. Write SQL to insert the example data into the enrollment tables.

Exercise 3.6. Add a few book records to the Book table without any authors. Then
execute the union and union all queries and observe the results.




                                                                                    55
4 List-Based Data Structures

Most programming languages provide some sort of support for storing collections of
similar elements. The most common way is to store elements in an array. That is,
elements are stored together in a contiguous chunk of memory and individual elements
are accessed using an index. An index represents an offset with respect to the first
element which is usually stored at index 0 (referred to as zero-indexing).
There are several disadvantages to using arrays, however. In particular, once allocated,
the capacity of an array is fixed. It cannot grow to accommodate new elements and
it cannot shrink if we end up removing elements. Moreover, we may not need the full
capacity of the array at any given point in a program, leading to wasted space. Though
libraries may provide convenience functions, in general all of the “bookkeeping” in an
array is up to us. If we remove an element in the middle of the array, our data may no
longer be contiguous. If we add an element in the array, we have to make sure to find an
available spot. Essentially, all of the organization of the array falls to the user.
A much better solution is to use a dynamic data structure called a List. A List is an
Abstract Data Type (ADT) that stores elements in an ordered manner. That is, there
is a notion of a “first” element, a “second” element, etc. This is not necessarily the
same thing as being sorted. A list containing the elements 10, 30, 5 is not sorted, but
it is ordered ( 10 is the first element, 30 is the second, and 5 is the third and final
element). In contrast to an array, the list automatically organizes the elements in some
underlying structure and provides an interface to the user that provides some set of core
functionality, including:
   • A way to add elements to the list
   • A way to retrieve elements from the list
   • A way to remove elements from the list
in some manner. We’ll examine the specifics later on, but the key aspect to a list is that,
in contrast to an array, it dynamically expands and contracts automatically as the user
adds/removes elements.
How a list supports this core functionality may vary among different implementations.
For example, the list’s interface may allow you to add an element to the beginning
of the list, or to the end of the list, or to add the new element at a particular index;
or any combination of these options. The retrieval of elements could be supported by
providing an index-based retrieval method or an iterator pattern that would allow a user



                                                                                        57
4 List-Based Data Structures

to conveniently iterate over every element in the list.
In addition, a list may provide secondary functionality as a convenience to users, making
the implementation more flexible. For example, it may be useful for a user to tell how
many elements are in the list; whether or not it is empty or full (if it is designed to have
a constrained capacity). A list might also provide batch methods to allow a user to add
a collection of elements to the list rather than just one at a time.
Most languages will provide a list implementation (or several) as part of their standard
library. In general, the best practice is to use the built-in implementations unless there
is a very good reason to “roll your own” and create your own implementation. However,
understanding how various implementations of lists work and what properties they provide
is very important. Different implementations provide advantages and disadvantages and
so using the correct one for your particular application may mean the difference between
an efficient algorithm and an inefficient or even infeasible one.



4.1 Array-Based Lists

Our first implementation is an obvious extension of basic arrays. We’ll still use a basic
array to store data, but we’ll build a data structure around it to implement a full list.
The details of how the list works will be encapsulated inside the list and users will interact
with the list through publicly available methods.
The basic idea is that the list will own an array (via composition) with a a certain
capacity. When users add elements to the list, they will be stored in the array. When the
array becomes full, the list will automatically reallocate a new, larger array (giving the
list a larger capacity), copy over all the elements in the old array and then switch to this
new array. We can also implement the opposite functionality and shrink the array if we
desire.


4.1.1 Designing a Java Implementation

To illustrate this design, consider the basic the following code sample in Java.

  1   public class IntegerArrayList {
  2

  3       private int arr[];
  4       private int size;
  5

  6   }


This array-based list is designed to store integers in the arr array. The second member



58
                                                                   4.1 Array-Based Lists

variable, size will be used to track the number of elements stored in arr . Note that
this is not the same thing as the size of the array (that is, arr.length ). Elements
may or may not be stored in each array position. To distinguish between the size of the
array-based list and the size of the internal array, we’ll refer to them as the size and
capacity
To initialize an empty array list, we’ll create a default constructor and instantiate the
array with an initial capacity of 10 elements with an initial size of 0 .

  1   public IntegerArrayList() {
  2     this.arr = new int[10];
  3     this.size = 0;
  4   }



Adding Elements

Now let’s design a method to provide a way to add elements. As a first attempt, let’s
allow users to add elements to the end of the list. That is, if the list currently contains
the elements 8, 6, 10 and the user adds the element 42 the list will then contain the
elements 8, 6, 10, 42 .
Since the size variable keeps track of number of elements in the list, we can use it to
determine the index at which the element should be inserted. Moreover, once we insert
the element, we’ll need to be sure to increment the size variable since we are increasing
the number of elements in the list. Before we do any of this, however, we need to check
to ensure that the underlying array has enough room to hold the new element and if it
doesn’t, we need to increase the capacity by creating a new array and copying over the
old elements. This is all illustrated in the following code snippet.

  1   public void addAtEnd(int x) {
  2

  3     //if the array is at capacity, resize it
  4     if(this.size == this.arr.length) {
  5       //create a new array with a larger capacity
  6       int newArr = new int[this.arr.length + 10];
  7       //copy over all the old elements
  8       for(int i=0; i<this.arr.length; i++) {
  9         newArr[i] = this.arr[i];
 10       }
 11       //use the new array
 12       this.arr = newArr;
 13     }
 14     this.arr[size] = x;



                                                                                        59
4 List-Based Data Structures


 15        this.size++;
 16    }


The astute Java programmer will note that lines 5–12 can be improved by utilizing
methods provided by Java’s Arrays class. Adhering to the Don’t Repeat Yourself
(DRY) principle, a better version would be as follows.

  1    public void addAtEnd(int x) {
  2

  3        if(this.size == this.arr.length) {
  4          this.arr = Arrays.copyOf(this.arr, this.arr.length + 10);
  5        }
  6        this.arr[size] = x;
  7        this.size++;
  8    }


As another variation, we could allow users to add elements at an arbitrary index. That
is, if the array contained the elements 8, 6, 10 , we could allow the user to insert the
element 42 at any index 0 through 3. The list would automatically shift elements down
to accommodate the new element. For example:
      • Adding at index 0 would result in 42, 8, 6, 10
      • Adding at index 1 would result in 8, 42, 6, 10
      • Adding at index 2 would result in 8, 6, 42, 10
      • Adding at index 3 would result in 8, 6, 10, 42
Note that though there is no element (initially) at index 3, we still allow the user to
“insert” at that index to allow the user to insert the element at the end. However, any
other index should be considered invalid as it would either be invalid (negative) or it
would mean that the data is no longer contiguous. For example, adding 42 at index
5 may result in 8, 6, 10, null, null, 42 . Thus, we do some basic index checking
and throw an exception for invalid indices.

  1    public void insertAtIndex(int x, int index) {
  2

  3        if(index < 0 || index > this.size) {
  4          throw new IndexOutOfBoundsException("invalid index: " + index);
  5        }
  6

  7        if(this.size == this.arr.length) {
  8          this.arr = Arrays.copyOf(this.arr, this.arr.length + 10);
  9        }



60
                                                                   4.1 Array-Based Lists

 10

 11       //start at the end; shift elements to the right to accommodate x
 12       for(int i=this.size-1; i>=index; i--) {
 13         this.arr[i+1] = this.arr[i];
 14       }
 15       this.arr[index] = element;
 16       this.size++;
 17   }


At this point we note that these two methods have a lot of code in common. In fact, one
could be implemented in terms of the other. Specifically, insertAtIndex is the more
general of the two and so addToEnd should use it:

  1   public void addAtEnd(int x) {
  2

  3       this.insertAtIndex(x, this.size);
  4   }


This is a big improvement as it reduces the complexity of our code and thus the complexity
of testing, maintenance, etc.


Retrieving Elements

In a similar manner, we can allow users to retrieve elements using an index-based retrieval
method. We will again take care to do index checking, but otherwise returning the
element is straightforward.

  1   public void getElement(int index) {
  2

  3       if(index < 0 || index >= this.size) {
  4         throw new IndexOutOfBoundsException("invalid index: " + index);
  5       }
  6       return this.arr[index];
  7   }



Removing Elements

When a user removes an element, we want to take care that all the elements remain
contiguous. If the array contains the elements 8, 6, 10 and the user removes 8 . we
want to ensure that the resulting list is 6, 10 and not null, 6, 10 . Likewise, we’ll
want to make sure to decrement the size when we remove an element. We’ll also return



                                                                                        61
4 List-Based Data Structures

the value that is being removed. The user is free to ignore it if they don’t want it, but
this makes the list interface a bit more flexible as it means that the user doesn’t have to
make two method calls to retrieve-then-delete the elements. This is a common idiom
with many collection data structures.

  1    public int removeAtIndex(int index) {
  2

  3        if(index < 0 || index > this.size) {
  4          throw new IndexOutOfBoundsException("invalid index: " + index);
  5        }
  6

  7        //start at index and shift elements to the left
  8        for(int i=index; i<size-1; i++) {
  9          this.arr[i] = this.arr[i+1];
 10        }
 11        this.size--;
 12    }


Note that we omitted automatically “shrinking” the underlying array if its unused
capacity became too big. We leave that and other variations on the core functionality as
an exercise.


Secondary Functionality

In addition to the core functionality of a list, you could extend our design to provide more
convenience methods to make the list implementation even more flexible. For example,
you may implement the following methods, the details of which are left as an exercise.

      • public boolean isEmpty() – this method would return true if no elements
        were stored in the list, false otherwise.
      • public int size() – more generally, a user may want to know how many ele-
        ments are in the list especially if they wanted to avoid an IndexOutOfBoundsException .

      • public void addAtBeginning(int element) – the companion to the addAtEnd(int)
        method
      • public int replaceElementAt(int element, int index) – a variation on the
        remove method that replaces rather than remove the element, returning the replaced
        element.
      • public void addAll(int arr[], int index) – a batch method that allows a
        user to add entire array to the list with one method call. Similarly, you could allow
        the user to add elements stored in another list instance,



62
                                                                   4.1 Array-Based Lists

       public void addAll(IntegerArrayList list, int index) . Sometimes this
      operation is referred to as “splicing.”
   • public void clear() – another batch method that allows a user to remove all
     elements at once.

Many other possible variations exist. In addition to the interface, one could vary how
the underlying array expands or shrinks to accommodate additions and deletions. In the
example above, we increased the size by a constant size of 10. Variations may include
expanding the array by a certain percentage or doubling it in size each time. Each
strategy has its own advantages and disadvantages.


A Better Implementation

The preceding list design was still very limited in that it only allowed the user to store
integers. If we wanted to design a list to hold floating point numbers, strings, or a
user defined type, we would need to create an implementation for every possible type
that we wanted a list for. Obviously this is not a good approach. Each implementation
would only differ in its name and the type of elements it stored in its array. This is a
quintessential example of when to use parameterized polymorphism. Instead of designing
a list that holds a particular type of element, we can parameterize it to hold any type of
element. Thus, only one array-based list implementation is needed.
In the example we designed before for integers, we never actually examined the content
of the array inside the class. We never used the fact that the underlying array held
integers and the only time we referred to the int type was in the method signatures
which can all be parameterized to accept and return the same type.
Code Sample 4.1 contains a parameterized version of the array-based list we implemented
before.
A few things to note about this parameterized implementation. First, with the original
implementation since we were storing primitive int elements, null was not an issue.
Now that we are using parameterized types, a user would be able to store null elements
in our list. We could make the design decision to allow this or disallow it (by adding null
pointer checks on the add/insert methods).
Another issue, particular to Java, is the instantiation of the array on line 7. We cannot
invoke the new keyword on an array with an indeterminate type. This is because
different types require a different number of bytes (integers take 4 bytes, double s take
8 bytes). The number of bytes may even vary between Java Virtual Machine (JVM)s
(32-bit vs. 64-bit). Without knowing how many bytes each element takes, it would be
impossible for the JVM to allocate the right amount of memory. Thus, we are forced to
use a raw type (the Object type) and do an explicit type cast. This is not much of an
issue because the parameterizations guarantee that the user would only ever be able to
add elements of type T .



                                                                                        63
4 List-Based Data Structures

Another useful feature that we could add would be an iterator pattern. An iterator allows
you to iterate over each element in a collection and process them. With a traditional
array, a simple for loop can be used to iterate over elements. An iterator pattern relieves
the user of the need to write such boilerplate code and instead use a foreach loop.1
In Java, an iterator pattern is achieved by implementing the Iterable<T> interface
(which is also parameterized). The interface requires the implantation of a public method
that returns an Iterator<T> which has several methods that need to be implemented.
The following is an example that could be included in our ArrayList implementation.

     1   public Iterator<T> iterator() {
     2     return new Iterator<T>() {
     3       private int currentIndex = 0;
     4       @Override
     5       public boolean hasNext() {
     6         return (this.currentIndex < size);
     7       }
     8

     9        @Override
 10           public T next() {
 11             this.currentIndex++;
 12             return arr[currentIndex-1];
 13           }
 14

 15          };
 16      }


Essentially, the iterator (an anonymous class declaration/definition) has an internal index,
 currentIndex that is initialized to 0 (the first element). Each time next() is called,
the “current” element is returned, but the method also sets itself up for the next iteration
by incrementing currentIndex . The main advantage of implementing an iterator is
that you can then use a foreach loop (which Java calls an “enhanced for-loop”). For
example:

     1   ArrayList<Integer> list = new ArrayList<Integer>();
     2   list.addAtEnd(8);
     3   list.addAtEnd(6);
     4   list.addAtEnd(10);
     5

     6   //prints "8 6 10"
 1
     Note that this is not necessarily mere syntactic sugar. As we will see later, some collections are
      unordered and would require the use of such a pattern. Yet still, some list implementations, in
      particular linked lists, an index-based get method is actually very inefficient. An iterator pattern
      allows you to encapsulate the most efficient logic for iterating over a particular list implementation.




64
                                                                           4.2 Linked Lists


  7   for(Integer x : list) {
  8     System.out.print(x + " ");
  9   }




4.2 Linked Lists

The array-based list implementation offers a lot of advantages and improvements over a
primitive array. However, it comes at some cost. In particular, when we need to expand
the underlying array, we have to create a new array and copy over every last element. If
there were 1 million elements in the underlying array and we wanted to add one more,
we would essentially be performing 1 million copy operations. In general, if there are n
elements, we would be performing n copy operations (or a copy operation proportional
to n). That is, some operations will induce a linear amount of work with respect to how
many elements are already in the list. Even with different strategies for increasing the
size of the underlying array (increasing the size by a percentage or doubling it), we still
have some built-in overhead cost associated with the basic list operations.
For many applications this cost is well-worth it. A copy operation can be performed
quite efficiently and the advantages that a list data structure provide to development
may outweigh the efficiency issues (and in any case, the same issues would be present
even if we used a primitive array). In some applications, however, an alternative
implementation may be more desirable. In particular, a linked list implementation avoids
the expansion/shrinking of an underlying array as it uses a series of linked nodes to store
elements rather than a primitive array.
A simple linked list is depicted in Figure 4.1. Each element is stored in a node. Each
node contains both an element (in this example, an integer) and a reference to the next
node in the list. In this way, each node is linked together to form a chain. The start
of the list is usually referred to as the head of the list. Similarly, the end of the list is
usually referred to as the tail (in this case, the node containing 10). A special symbol
or value is usually used to denote the end of the list so that that tail node does not
point to another node. In this case, we’ve used the value φ to indicate the end of the
list. In practice, a null value could be used or a special sentinel node can be created
to indicate the end of a list.
To understand how a linked list works, we’ll design several algorithms to support the
core functionality of a list data structure. In order to refer to nodes and their elements,
we’ll use the following notation. Suppose that u is a node, then its value will be referred
to as u.value and the next node it is linked to will be referred to as u.next.



                                                                                          65
4 List-Based Data Structures

                        head


                         8             6             10              φ

                      Figure 4.1: A simple linked list containing 3 nodes.

Adding Elements

With a linked list, there is no underlying array of a fixed size. To add an element, we
simply need to create a new node and link it somewhere in the chain. Since there is no
fixed array space to fill up, we no longer have to worry about expanding/shrinking it to
accommodate new elements.
To start, an empty list will be represented by having the head reference refer to the
special end-of-list symbol. That is, head → φ. To add a new element to an empty list,
we simply create a new node and have the head reference point to it. In fact, we can
more generally support an insert at head operation with the same idea. To insert at the
head of the list, we create a new node containing the inserted value, make it point to the
“old head” node of the list and then update the head reference to the new node. This
operation is depicted in Algorithm 1.

         Input    : A linked list L with head, L.head and a new element to insert x at the
                    head
     1   u ← a new node
     2   u.value ← x
     3   u.next ← L.head
     4   L.head ← u

                 Algorithm 1: Insert-At-Head Linked List Operation

Just as with the array-based list, we could also support a more general insert-at-index
method that would allow the user to insert a new node at any position in the list. The
first step, of course, would be to find the two nodes between which you wanted to insert
the new node. We will save the details for this procedure as it represents a more general
retrieval method. For now, suppose we have two nodes, a, b and we wish to insert a new
node, u between them. To do this, we simply need to make a refer to the new node
and to make the new node refer to b. However, we must be careful with the order in
which we do this so as not to lose a’s reference to b. The general procedure is depicted
in Algorithm 2.
One special corner case occurs when we want to insert at the end of a list. In this



66
                                                                                       4.2 Linked Lists




               head                                                        head

               8              6                               42            8           6


      (a) Initially the linked list has element 8 as its     (b) We create a new node containing 42
      head.

                      head                                   head

      42              8             6                        42            8           6


  (c) We then make the new node refer to the old           (d) And finally update the head reference to
  head node.                                               point to the new node instead.

Figure 4.2: Insert-at-head Operation in a Linked List. We wish to insert a new element,
            42 at the head of the list.




      Input    : Two linked nodes, a, b in a linked list and a new value, x to insert
                 between them
  1   u ← a new node
  2   u.value ← x
  3   u.next ← a.next
  4   a.next ← u

                   Algorithm 2: Insert Between Two Nodes Operation




                                                                                                          67
4 List-Based Data Structures



              8                6                                     8                 6


                      42                                                     42
      (a) We create a new node containing 42           (b) We then make the new node point to the second
                                                       node



              8                6
                                                                8            42            6
                      42
(c) And reassign the first node’s next reference to the (d) Resulting in the new node being inserted between
new node.                                               the two given nodes.

Figure 4.3: Inserting Between Two Nodes in a Linked List. Here, we wish to insert a
            new element 42 between the given two nodes containing 8 and 6.


scenario, b would not exist and so a.next would end up referring to φ. This ends up
working out with the algorithm presented. We never actually made any explicit reference
to b in Algorithm 2. When we assigned u.next to refer to a.next, we took care of both
the possibility that it was an actual node and the possibility that it was φ.
Another corner case occurs if we wish to insert at the head of the list using this algorithm.
In that scenario, a would not refer to an actual node while b would refer to the head
element. In this case, lines 3–4 would be invalid as a.next would be an invalid reference.
For this corner case, we would need to either fall back to our first insert-at-head (Algorithm
1) operation or we would need to handle it separately.


Retrieving Elements

As previously noted, a linked list avoids the cost of expanding/shrinking of an underlying
array. However, this does come at a cost. With an array-based list, we had “free” random
access to elements. That is, if we wanted the element stored at index i it is a simple
matter to compute a memory offset and “jump” to the proper memory location. With a
linked list, we do not have the advantages of random access.
Instead, to retrieve an element, we must sequentially search through the list, starting from
the head, until we find the element that we are trying to retrieve. Again, many variations
exist, but we’ll illustrate the basic functionality by describing the same index-based
retrieval method as before.



68
                                                                          4.2 Linked Lists


      Input   : A linked list L with head, L.head and in index i, 0 ≤ i < n where n is
                the number of elements in L
      Output : The i-th node in L
  1   currentN ode ← L.head
  2   currentIndex ← 0
  3   while currentIndex < i do
  4      currentN ode ← currentN ode.next
  5      currentIndex ← (currentIndex + 1)
  6   end
  7   output currentN ode

                 Algorithm 3: Index-Based Retrieval Operation


The key to this algorithm is to simply keep track of the “current” node. Each iteration
we traverse to the next node in the chain and iterate a counter. When we have traversed
i times, we stop as that is the node we are looking for.
In contrast to the “free” index-based retrieval method with an array-based list, the
operation of finding the i-th node in a linked list is much more expensive. We have to
perform i operations to get the i-th node. In the worst case, when we are trying to
retrieve the last (tail) element, we would end up performing n traversal operations. If
the linked list held a lot of elements, say 1 million, then this could be quite expensive.
In this manner, we see some clear trade-offs in the two implementations.


Removing Elements

Like the insertion operation, the removal of an element begins by retrieving the node
that contains it. Suppose we have found a node, u and wish to remove it. It actually
suffices to simply circumvent the node, by making u’s predecessor point to u’s successor.
This is illustrated in Figure 4.4.
Since we are changing a reference in u’s predecessor, however, we must make appropriate
changes to the retrieval operation we developed before. In particular, we must keep track
of two nodes: the current node as before but also its predecessor, or more generally, a
previous node. Alternatively, if we were performing an index-based removal operation,
we could easily find the predecessor by adjusting our index. If we wanted to delete the
i-th node, we could retrieve the (i − 1)-th node and delete the next one in the list.
Again, we may need to deal with corner cases such as if we are deleting the head of the
list or the tail of the list. As with the insertion, the case in which we delete the tail is
essentially the same as the general case. If the successor of a tail node is φ, thus when
we make u’s predecessor refer to u’s successor, we are simply making it refer to φ.



                                                                                         69
4 List-Based Data Structures




                                                         Node to delete



                        8                6                10


                       (a) We wish to delete the node containing 6.




                        8                6                10


                  (b) We make its predecessor node point to its successor.




                                8                10


                     (c) Though the node containing 6 still refers to the
                     next node, from the perspective of the list, it has
                     been removed.

                   Figure 4.4: Delete Operation in a Linked List




70
                                                                        4.2 Linked Lists

The more complicated part is when we delete the head of the list. By definition, there is
no predecessor to the head, so handling it as the general case will result in an invalid
reference. Instead, if we wanted to delete the head, we must actually change the list’s
head itself. This is easy enough, we simply need to make head refer to head− > next.
An example of a removal operation is presented in Algorithm 4. In contrast to previous
examples, this operation is a key-based removal operation variation. That is, we search
for the first instance of a node whose value matches a given key element and delete it.
As another corner case for this variation, we also must deal with the situation in which
no node matches the given key. In this case, we’ve decided to leave it as a no-operation
(“noop”). This is achieved in lines 10–12 which will not delete the last node if the key
does not match.

      Input : A linked list L with head, L.head and a key element k
      Output : L but with the first node u such that u.value = k removed if one exists
  1   if L.head.value = k then
  2       L.head. ← L.head.next
  3   else
  4       previousN ode ← φ
  5       currentN ode ← L.head
  6       while currentN ode.value 6= k and currentN ode.next 6= φ do
  7          previousN ode ← currentN ode
  8          currentN ode ← currentN ode.next
  9      end
 10      if currentN ode.value = k then
 11          previousN ode.next ← currentN ode.next
 12      end
 13   end
 14   output L

                   Algorithm 4: Key-Based Delete Operation



4.2.1 Designing a Java Implementation

We now adapt these operations and algorithms to design a Java implementation of
a linked list. Note that the standard collections library has both an array-based list
( java.util.ArrayList<E> ) as well as a linked list implementation ( java.util.LinkedList<E> )
that should be used in general.
First, we need a node class in order to hold elements as well as a reference to another
node. Code Sample 4.2 gives a basic implementation with several convenience methods.



                                                                                         71
4 List-Based Data Structures




      φ            8              6               10              42                7


  previous       current
                       (a) Initialization of the previous and current references.

                  8               6              10               42                7


                previous        current
          (b) After the first iteration of the while loop, the references have been updated.

                  8               6              10               42                7


                                previous        current
                           (c) After the second iteration of the while loop.

                  8               6              10               42                7


                                                previous         current
(d) After the third iteration, the current node matches our key value, 42 and the loop terminates.




                  8               6              10               42                7


                                                previous         current
           (e) Since the current node matches the key, we remove it by circumventing it.



                  8               6              10               7
                              (f) Resulting in the node’s removal.

Figure 4.5: Key-Based Find and Remove Operation. We wish to remove the first node
            we find containing 42.




72
                                                                         4.2 Linked Lists

The class is parameterized so that nodes can hold any type.
Given this Node class, we can now define the basic state of our linked list implementation:

  1   public class LinkedList<T> {
  2

  3       private Node<T> head;
  4       private int size;
  5

  6       public LinkedList() {
  7         this.head = null;
  8         this.size = 0;
  9       }
 10

 11       ...
 12

 13   }


We keep track of the size of the list and increment/decrement it on each add/delete
operation so that we do not have to recompute it. To keep things simple, we will
implement two general purpose node retrieval methods, one index-based and one key
based. These can be used as basic steps in other, more specific operations such as
insertion, retrieval, and deletion methods. Note that both of these methods are private
as they are intended for “internal” use by the class and not for external use. If we had
made these methods public we would be exposing the internal structure of our linked
list to outside code. Such a design would be a typical example of a leaky abstraction and
is, in general, considered bad practice and bad design.

  1   private Node<T> getNodeAtIndex(int index) {
  2

  3       if(index < 0 || index >= this.size) {
  4         throw new IndexOutOfBoundsException("invalid index: " + index);
  5       }
  6       Node<T> curr = this.head;
  7       for(int i=0; i<index; i++) {
  8         curr = curr.getNext();
  9       }
 10       return curr;
 11

 12   }
 13

 14   private Node<T> getNodeWithValue(T key) {
 15

 16       Node<T> curr = head;



                                                                                        73
4 List-Based Data Structures


 17       while(curr != null && !curr.getItem().equals(key)) {
 18         curr = curr.getNext();
 19       }
 20       return curr;
 21   }



As previously mentioned, these two general purpose methods can be used or adapted
to implement other, more specific operations. For example, index-based retrieval and
removal methods become very easy to implement using these methods as subroutines.


  1   public T getElement(int index) {
  2     return getNodeAtIndex(index).getItem();
  3   }
  4

  5   public T removeAtIndex(int index) {
  6

  7       if(index < 0 || index >= this.size) {
  8         throw new IndexOutOfBoundsException("invalid index: " + index);
  9       }
 10       Node<T> previous = getNodeAtIndex(index-1);
 11       Node<T> current = previous.getNext();
 12       T removedItem = current.getItem();
 13       previous.setNext(current.getNext());
 14       return removedItem;
 15   }



Adapting these methods and using them to implement other operations is left as an
exercise.




4.2.2 Variations

In addition to the basic linked list data structure design, there are several variations
that have different properties that may be useful in various applications. For example,
one simple variation would be to not only keep track of the head element, but also a
tail element. This would allow you to efficiently add elements to either end of the list
without having to traverse the entire list. Other variations are more substantial and we
will not look at several of them.



74
                                                                          4.2 Linked Lists

                      head                                    tail


         φ                8                 6                  10               φ

                       Figure 4.6: A Doubly Linked List Example


Doubly Linked Lists

As presented, a linked list is “one-way.” That is, we can only traverse forward in the list
from the head toward the tail. This is because our tree nodes only kept track of a next
element, referring to the node’s predecessor. As an alternative, our list nodes could keep
track of both the next element as well as a previous element so that it has access to both
a node’s predecessor as well as its successor. This allows us to traverse the list two ways,
forwards and backwards. This variation is referred to as a doubly linked list.
An example of a doubly linked list is depicted in Figure 4.6. In this example, we’ve also
established a reference to the tail element to enable us to start at the end of the list and
traverse backwards.
A doubly linked list may provide several advantages over a singly linked list. For example,
when we designed our key-based delete operation we had to be sure to keep track of a
previous element in order to manipulate its reference to the next node. In a doubly linked
list, however, since we can always traverse to the previous node this is not necessary.
A doubly linked list has the potential to simplify a lot of algorithms, however it also
means that we have to take greater care when we manipulate the next and previous
references. For example, to insert a new node, we would have to modify up to four
references rather than just two. Greater care must also be taken with respect to corner
cases.


Circularly Linked Lists

Another variation is a circularly linked list. Instead of a ending the list using a sentinel
value, φ, the “last” node points back to the first node, closing the list into a large loop.
An example is given in Figure 4.7. With such a structure, it is less clear that there is a
head or tail to the list. Certain operations such as index-based operations may no longer
make sense with such an implementation. However, we could still designate an arbitrary
node in the list as the “head” as in the figure.
Alternatively, we could instead simply have a reference to a current node. At any point
during the data structure’s life cycle the current node may reference any node in the list.
The core operations may iterate through the list and end up at a different node each



                                                                                         75
4 List-Based Data Structures

                          head


                          8             6               10


                     Figure 4.7: A Circularly Linked List Example

                                      m elements

                head            81    16    ···    10    32


                                21    42    ···    89    47


                                13    17    ···    21    18           φ

                     Figure 4.8: An Unrolled Linked List Example


time. Care would have to be taken to ensure that operations such as searching do not
result in an infinite loop, however. An unsuccessful key-based search operation would
need to know when to terminate (when it has gone through the entire circle once and
returned to where it started). It is easy enough to keep track of the size of the list and
to ensure that no circular operations exceed this size.
Circularly linked lists are useful for applications where elements must be processed over
and over each in turn. For example, an operating system may give time slices to running
applications. Instead of a well-defined beginning and end, a continuous poll loop is run.
When the “last” application has exhausted its time slot, the operating system returns to
the “first” in a continuous loop.


Unrolled Linked Lists

As presented, each node in a linked list holds a single element. However, we could design
a node to hold any number of elements, in particular an array of m elements. Nodes
would still be linked together, but each node would be a mini array-based list as well.
An example is presented in Figure 4.8.



76
                                                                       4.3 Stacks & Queues

This hybrid approach is intended to achieve better performance with respect to memory.
The size of each node may be designed to be limited to fit in a particular cache line, the
amount of data typically transferred from main memory to a processor’s cache. The goal
is to improve cache performance while reducing the overhead of a typical linked list. In a
typical linked list, every node has a reference (or two in the case of doubly linked lists)
which can double the amount of memory required to store elements. By storing more
elements in each node, the overall number of nodes is reduced and thus the number of
references is reduced. At the same time, an unrolled linked list does not require large
chunks of contiguous storage like an array-based list but provides some of the advantages
(such as random access within a node). Overall, the goal of an unrolled linked list is to
reduce the number of cache misses required to retrieve a particular element.


4.3 Stacks & Queues

Collection data structures such as lists are typically unstructured. A list, whether array-
based, a linked list, or some other variation simply hold elements in an ordered manner.
That is, there is a notion of a first element, second element, etc. However that ordering
is not necessarily structured, it is simply the order in which the elements were added to
the collection. In contrast, you can impose a structured ordering by sorting a list (thus,
“sorted” is not the same thing as “ordered”). Sorting a list, however, does not give the
data structure itself any more structure. The interface would still allow a user to insert
or rearrange the elements so that they are no longer sorted. Sorting a list only change’s
the collection’s state, not its behavior.
We now turn our attention to a different kind of collection data structure whose structure
is defined by its behavior rather than its state. In particular, we will look at two data
structures, stack and queues, that represent restricted access data structures. These data
structures still store elements, but the general core functionality of a list (the ability
to add, retrieve, and remove arbitrary elements) will be restricted in a very particular
manner so as to give the data structure behavioral properties. This restriction is built
into the data structure as part of the object’s interface. That is, users may only interact
with the collection in very specific ways. In this manner, structure is imposed through
the collection’s behavior.


4.3.1 Stacks

A stack is a data structure that stores elements in a last-in, first-out (or Last-In First-Out
(LIFO) manner. That is, the last element to be inserted into a stack is the first element
that will come out of the stack. A stack data structure can be described as a stack of
dishes. When dealing with such a stack, we can add a dish to it, but only at the top of
the stack lest we risk causing the entire stack of dish to fall and break. Likewise, when
removing a dish, we remove the top-most dish rather than pulling a dish from the middle



                                                                                           77
4 List-Based Data Structures

or bottom of the stack for the same reason. Thus, the last dish that we added to the
stack will be the first dish that we take off the stack. It may also be helpful to visualize
such a stack of dishes as in a cafeteria where a spring-loaded cart holds the stack. When
we add a dish, the entire stack moves down into the cart and when we remove one, the
next one “pops” up to the top of the stack.
You are probably already familiar with the concept of stacks in the context of a program’s
call stack. As a program invokes functions, a new stack frame is created that contains all
the “local” information (parameters, local variables, etc.) and is placed on top of the call
stack. When a function is done executing and returns control back to the calling function
the stack frame is removed from the top of the call stack, restoring the stack frame right
below it. In this manner, information can be saved and restored from function call to
function call efficiently. This idea goes all the way back to the very earliest computers
and programming languages (specifically, the Information Processing Language in 1956).


Core Functionality

In order to have achieve the LIFO behavior, access to a stack’s elements are restricted
through its interface by only allowing two core operations:
     • push adds a new element to the stop of the stack and
     • pop removes the element at the top of the stack
The element removed at the top of the stack may also be “returned” in the context of a
method call.
Similar to lists, stacks also have several corner cases that need to be considered. First,
when implementing a stack you must consider what will happen when a user performs
a pop operation on an empty stack. Obviously there is nothing to remove from the
stack, but how might we handle this situation? In general, such an invalid operation is
referred to as a stack underflow. In programming languages that support error handling
via exceptions, we could choose to throw an exception. If we were to make this design
decisions then we should, in addition, provide a way for the user to check that such an
operation may result in an exception by providing a way for them to check if the stack is
empty or not (see Secondary Functionality below). Alternatively, we could instead return
a “flag” value such as null to indicate an empty stack. Though this design decision has
consequences as well: we would either need to disallow the user from pushing a null
value onto the stack (and decide how again to handle that) or we would need to provide
a way for the user to distinguish the situation where null was actually popped off the
stack or was returned because the stack was empty.
In addition, we could design our stack to be either bounded or unbounded. An unbounded
stack means that there would be no practical restrictions on how large the stack could
grow. A program’s use of our stack would only be limited by the amount of system
memory available. A user could continue to push as many elements onto the stack as



78
                                                                   4.3 Stacks & Queues




                            push                     pop



                                                   Top

                                           7
                                           3
                                          42
                                          10
                                           6
                                           8       Bottom

Figure 4.9: A stack holding integer elements. Push and pop operations are depicted
            as happening at the “top” of the stack. In actuality, a stack stored in a
            computer’s memory is not really oriented but this visualization is consistent
            with a physical stack growing “upwards.”




                                                                                      79
4 List-Based Data Structures

they like and a problem would only arise when the program or system itself runs out of
memory.
A bounded stack means that we could design our stack to have a fixed capacity or limit
of (say) n elements. If we went with such a design we would again have to deal with
the corner case of a user pushing an element to the stack when it is full referred to as
a stack overflow. Solutions similar to the popping from an empty stack could be used
here as well. We could throw an exception (and give the user the ability to check if the
stack is full or not) or we could make such an operation a no-op: we would not push the
element to the stack, leaving it as it was before and then report the no operation to the
user. Typically a boolean value is used, true to indicate that the operation was valid
and had some side effect on the stack (the element was added) or false to indicate that
the operation resulted in no side effects.


Secondary Functionality

To make a stack more versatile it is common to include secondary functionality such as
the following.

     • A peek method that allows a user to access the element at the top of the stack
       without removing it. The same effect can be achieved with a pop-then-push
       operation, but it may be more convenient to allow such access directly. This is
       useful if a particular algorithm needs to make a decision based on what will be
       popped off the stack next.
     • A means to iterate over all the elements in the stack or to allow read-only access
       arbitrary elements. We would not want to allow arbitrary write access as that
       would violate the LIFO behavior of a stack and defeat the purpose of using this
       particular data structure.
     • A way to determine how many elements are on the stack and, related whether
       or not the stack is empty and, if it is bounded, whether or not it is full or its
       remaining capacity. Such methods would allow the user to use a more defensive-style
       programming approach and not make invalid operations.
     • A way to empty or “clear” the stack of all its elements with a single method call.
     • General find or contains methods that would allow the user to determine if a
       particular element was already in the stack and more generally, if it is, how far
       down in the stack it is (its “index”).


Implementations

A straightforward and efficient implementation for a stack is to simply use a list data
structure “under the hood” and to restrict access to it through the stack’s interface. The



80
                                                                     4.3 Stacks & Queues

push and pop operations can then be achieved in terms of the list’s add and remove
operations, taking care that both work from the same “end” of the list. That is, if the
push operation adds an element at the beginning of the list, then the pop operation must
remove (and return) the element at the beginning of the list as well.
A linked list is ideal for bounded and unbounded stacks as adding and removing from
the head of the list are both very efficient operations, requiring only the creation of
a new node and the shuffling of a couple of references. There is also no expensive
copy-and-expand operation over the life of the stack. For unbounded stacks, the capacity
can be constrained by simply checking the size of the underlying list and handling the
stack overflow appropriately. If the underlying linked list also keeps track of the tail
element, adding and removing from the tail would also be an option.
An array-based list may also be a good choice for bounded stacks if the list is initialized
to have a capacity equal to the capacity of the stack so as to avoid any copy-and-expand
operations. An array-based list is less than ideal for unbounded stacks as expensive
copy-and-expand operations may be common. However, care must be taken to ensure
that the push and pop operations are efficient. If we designate the “first” element (the
element at index 0) as the top of the stack, we would constantly be shifting elements
with each and every push and pop operation which can be quite expensive. Instead, it
would be more efficient to keep track of the top element and add elements to the “end”
of the array.
In detail, we would keep track of a current index, say top that is initialized to −1
indicating an empty stack. Then as elements are pushed, we would add them to the list
at index (top + 1) and increment top. As elements are popped, we return the element
at index top and decrement the index. In this manner, each push and pop operation
requires a constant number of operations rather than shifting up to n elements.


Applications

As previously mentioned, stacks are used extensively in computer architecture. They are
used to keep track of local variables and parameters as functions are called using a program
stack. In-memory stacks may also be used to simulate a series of function/method calls
in order to avoid using (or misusing) the program stack. A prime example of such a use
case is avoiding recursion. Instead of a sequence of function calls that may result in a
stack overflow of the program stack (which is generally limited) an in-memory stack data
structure, which is generally able to accommodate many more elements, can be used.
Stacks are also the core data structures used in many fundamental algorithms. Stacks
are used extensively in algorithms related to parsing and processing of data such as in
the Shunting Yard Algorithm used by compilers to evaluate an Abstract Syntax Tree
(AST). Stacks are also used in many graph algorithms such as Depth First Search (DFS)
and in a preorder processing of binary trees (see Chapter 6). Essentially any application
in which something needs to be “tracked” or remembered in a particular LIFO ordering,



                                                                                         81
4 List-Based Data Structures

       dequeue



                         8     6      10     42     3      7

                    Front                                      End
                                                                           enqueue

Figure 4.10: An example of a queue. Elements are enqueued at the end of the queue and
             dequeued from the front of the queue.


a stack is an ideal data structure to use.



4.3.2 Queues

A similar data structure is a queue which provides a First-In First-Out (FIFO) ordering
of elements. As its name suggests, a queue can be thought of as a line. Elements enter
the queue at one end (the “end” of the line) and are removed from the other end (the
“front” of the line). Thus, the first element to enter a queue is the first element to be
removed from the queue. Each element in the “line” is served in the order in which they
entered the line. Similar to a stack, a queue’s structure is defined by its interface.



Core Functionality

The two core operations in a queue are:
     • enqueue which adds an element to the queue at its end and
     • dequeue which removes the element at the front of the queue
An example of a queue and its operations is depicted in Figure 4.10. Some programming
languages and data structure implementations may use different terminology to describe
these two operations. Some use the same terms as a stack (“push” would correspond
to enqueue and “pop” would correspond to a dequeue operation). However, using
the same terminology for fundamentally different data structures is confusing.2 Some
implementations use the terms “offer” (we are offering an element to the queue if it is
able to handle it) and “poll” (we are asking or “polling” the queue to see if it has any
elements to remove).

 2
     <opinion>and wrong</opinion>




82
                                                                       4.3 Stacks & Queues

Secondary Functionality

Secondary functionality is pretty much the same as with a stack. We could choose to
make our queue bounded or unbounded and deal with corner cases (enqueuing to a full
queue, dequeueing from an empty queue) similarly. In fact, we would probably want to
design our queue with the same behavior for consistency. We could include methods to
determine the size of the queue, its remaining capacity (if bounded), a way to determine
if (and where) a certain element may be in the queue, etc.
As with stacks, in general we would not want to allow a user to arbitrarily insert elements
into a queue. Doing so would be allowing “line jumpers” to jump ahead of other elements,
violating FIFO. In some applications this does make sense and we discuss these variations
in Section 4.3.3. However, we could allow arbitrary removal of certain elements. Strictly
speaking, this would not violate FIFO as the remaining elements would still be processed
in the order in which they were enqueued. This could model situations where those
waiting in line got impatient and left (a process or request timed-out for example).


Implementations

The obvious implementation for a queue is a linked list. Since we have to work from
both ends, however, our linked list implementation will need to keep track of both the
head element and the tail element so that adding a new node at either end has the same
constant cost (creating a new node, shuffling some references). If our linked list only
keeps track of the head, then to enqueue an element, we would need to traverse the entire
list all the way to the end in order to add an element to the tail.
A linked list that offers constant-time add and remove methods to both the head and
the tail can be oriented either way. We could add to the head and remove from the tail
or we could add to the tail and remove from the head. As long as we are consistently
adding to one end and removing from the other, there really is no difference. Our design
decision may have consequences, however, on the secondary functionality. If we design
the array with an iterator, it makes the most sense to start at the front of the queue and
iterate toward the end. If our linked list is a doubly linked list, then we can easily iterate
in either direction. However, if our list is singly linked then we need to make sure that
the front of our queue corresponds to the head of the list.
An array-based list can also be used to implement a queue. As with stacks, it is most
appropriate to use array-based lists when you want a bounded queue so as to avoid
expensive expand-and-copy operations. However, you need to be a bit clever in how you
enqueue and dequeue items to the array in order to ensure efficient operations.
A naive approach would be to enqueue elements at one end of the array and dequeue
them from the front (index 0). However, this would mean we need to shift elements down
on every dequeue operation which is potentially very inefficient. A clever workaround
would be to keep track of the f ront and end of the queue using two index variables.



                                                                                           83
4 List-Based Data Structures

Initially, an empty queue would have both of these variables initialized to 0. As we add
elements, the end index gets incremented (we add left-to-right). As elements are removed,
the f ront index variable gets incremented. At some point, these variables will read the
right-end of the array at which point we simply reset them back to zero. The queue is
empty when f ront = end and it is full when (end − f ront) mod n = n − 1 where n is
the size of the array. The various states of such a queue are depicted in Figure 4.11.
Using an array-based list may save a substantial amount of memory. However, the added
complexity of implementation (and thus increased opportunities for bugs and errors) may
not justify the savings.




Applications


A queue is ideal for any application in which elements must be stored in order to be
handled or processed in a particular order. For example, a queue is a natural data structure
to implement buffers in which data is received but cannot be processed immediately (it
may not be possible or it would be inefficient to process the data in small chunks). A
queue ensures that the data remains in the order in which it was received.
Another typical example is when requests are received and must be processed. This is
typical in a webserver for example where requests for resources or webpages are stored in
a queue and processed in the order in which they are received. In an operating system,
threads (or processes) may be stored in a job queue and then woken/executed.
More generally, queues can facilitate communication in a Producer Consumer Pattern (see
Figure 4.12). In this scenario we have independent producers and consumers. Producers
may produce requests, tasks, works, or a resource that needs to be processed. For
example, a producer may be a web browser requesting a particular web page, or it may be
a thread making a request for a resource, etc. Consumers handle or service each of these
requests. Each consumer and each producer acts independently and asynchronously. To
facilitate thread-safe communication between the two groups, each request is enqueued
to a blocking queue. As producers enqueue requests, they are stored in the order they
are received. Then, as each consumer becomes available to service a request, it polls the
queue for the next request.
A naive approach would be to do busy polling where a consumer keeps polling the queue
over and over for another request even if it is empty. Using a thread-safe blocking queue
means that we don’t do busy polling. Instead, if no request is available for a consumer,
the queue blocks the consumer (typically, the consumer is a thread and this puts the
thread to sleep) until a request becomes available so that it is not continuously polling
for more work.



84
                                                                                     4.3 Stacks & Queues




           0       1      2       3    4      5            0        1       2        3         4       5

           –       –      –       –    –      –            8        6       10       42        –       –


       f ront                                         f ront                     end
            end
     (a) An initially empty queue. Both index       (b) As elements are enqueued the end index
     variables refer to index 0.                    variable is incremented.



          0       1      2       3    4       5       0         1       2        3         4       5

          –       –     10      42    3       7       67       13       –        –         3       7


                       f ront                end               end                       f ront

       (c) More elements may be enqueued as        (d) At some point the index variables wrap
       well as dequeued, moving both index vari-   around to the beginning of the array and
       ables.                                      the queue may “straddle” the array.



          0       1      2       3    4       5       0         1       2        3         4       5

          67      13    17      90    3       7       –         –       –        –         –       –


                                end f ront                                  f ront
                                                                                 end
       (e) The queue may also become full, at      (f) The queue may again become empty
       which point the two index variables are     and the two index variables refer to the
       beside each other.                          same value, but not necessarily index 0.

Figure 4.11: Various states of an array-based queue. As the index variables, f ront and
             end become n or larger, they wrap around to the beginning of the array
             using a modulus operation.




                                                                                                           85
4 List-Based Data Structures

         Consumers                                                       Producers
                                  waiting requests
                                                                               p
     c                                                                                   p
               c
                             r0      r1     r2       r3

         c                                  Blocking Queue                           p
                   c
                                                                          p

Figure 4.12: Producer Consumer Pattern. Requests (or tasks or resources) are enqueued
             into a blocking (or otherwise thread safe queue) by producers. Independently
             (asynchronously) consumers handle requests. Once done, consumers poll the
             queue for another request. If the queue is empty it blocks consumers until a
             new request becomes available.




4.3.3 Variations


In addition to the fundamental stack and queue data structures, there are numerous
useful variations. Some simple variations involve allowing the user to perform more than
just the two core operations on each.
For example, we could design a double-ended stack which, in addition to the push and
pop operations at the top, we could allow a pop operation at the bottom. This provides
a variation on the usual bounded stack where the “oldest” elements at the bottom drop
out when the stack becomes full instead of rejecting the “newest” items at the top of
the queue. A prime example of this is how a typical “undo” operation is facilitated in
many applications. Take for example a word processor in which the user performs many
actions in sequence (typing, highlighting, copy-paste, delete, etc.). A word processor
typically allows the user to undo previous operations but in the reverse sequence that
they were performed. However, at the same time we don’t want to keep track of every
change as it would start to take more and more memory and may impact performance.
Using a double-ended stack means that the oldest action performed is removed at the
bottom, allowing the last n operations to be tracked.
Relatedly, we could design a queue in which we are allowed to enqueue elements at either
end, but only remove from one. This would allow the aforementioned “line jumpers” to
jump ahead to the front of the line. This establishes a sort-of “fast lane” for high priority
elements. As an example, consider system processes in an operating system waiting for
their time slice to execute. User processes may be preempted by system-level processes
as they have a high priority. More generally, we can establish more than two levels of
priority (high/low) and build what is known as a priority queue (see below).



86
                                                                   4.3 Stacks & Queues

Deques

A logical extension is to allow insert and remove operations at both ends. Such a
generalized data structure is known as a deque (pronounced “deck”, also called a double-
ended queue). In fact, the primary advantage to creating this generalized data structure
is that it can be used as a stack, queue, etc. all with only a single implementation.
In fact in Java (as of version 6), it is recommended to use its java.util.Deque<E>
interface rather than the stack and queue implementations. The Deque<E> interface is
extremely versatile, offering several different versions of each of the insert and remove
methods to both ends (head and tail). One set of operations will throw exceptions
for corner cases (inserting into a full deque and removing from an empty deque) and
another set will return special values ( null or false ). Java also provides several
different implementations including an array-based deque ( ArrayDeque<E> ) and its
usual LinkedList<E> implementation as well as several concurrent and thread-safe
versions.


Priority Queues

Another useful variation is a priority queue in which elements are stored not in a
FIFO manner, but with respect to some priority. When elements are dequeued, the
highest priority element is removed from the queue first. When elements are enqueued,
conceptually they are placed in the queue according to their priority. This may mean
that the new element jumps ahead to the front of the queue (if it has a priority higher
than all other elements) or it may mean it ends up at the end of the queue (if it has the
lowest priority) or somewhere in between. In general, any scheme can be used to define
priority but it is typical to use integer values.
A naive implementation of a priority queue would implement the enqueue operation by
making comparisons and inserting the new element at the appropriate spot, requiring
up to n comparisons/operations in a queue with n elements. There are much better
implementations that we’ll look at later on (in particular, a heap implementation, see
Chapter 6. Again, most programming languages will have a built-in implementation.
Java provides an efficient heap-based PriorityQueue<E> implementation. Priority is
defined using either a natural ordering or a custom Comparator object.




                                                                                      87
4 List-Based Data Structures


  1   public class ArrayList<T> {
  2
  3       private T[] arr;
  4       private int size;
  5
  6       public ArrayList() {
  7         this.arr = (T[]) new Object[10];
  8         this.size = 0;
  9       }
 10
 11       public T getElement(int index) {
 12

 13           if(index < 0 || index >= this.size) {
 14             throw new IndexOutOfBoundsException("invalid index: " + index);
 15           }
 16           return this.arr[index];
 17       }
 18

 19       public void removeAtIndex(int index) {
 20
 21           if(index < 0 || index >= size) {
 22             throw new IndexOutOfBoundsException("invalid index: " + index);
 23           }
 24
 25           for(int i=index; i<size-1; i++) {
 26             this.arr[i] = this.arr[i+1];
 27           }
 28           this.size--;
 29       }
 30
 31       public void insertAtIndex(T x, int index) {
 32         if(index < 0 || index > size) {
 33           throw new IndexOutOfBoundsException("invalid index: " + index);
 34         }
 35

 36           if(this.size == arr.length) {
 37             this.arr = Arrays.copyOf(this.arr, this.arr.length + 10);
 38           }
 39
 40           for(int i=this.size-1; i>=index; i--) {
 41             this.arr[i+1] = this.arr[i];
 42           }
 43           this.arr[index] = element;
 44           this.size++;
 45       }
 46

 47       public void addAtEnd(T x) {
 48         this.insertAtIndex(x, this.size);
 49       }
 50   }


                   Code Sample 4.1: Parameterized Array-Based List in Java


88
                                                                4.3 Stacks & Queues




  1   public class Node<T> {
  2

  3       private final T item;
  4       private Node<T> next;
  5

  6       public Node(T item) {
  7         this.item = item;
  8         next = null;
  9       }
 10

 11       //getters and setters omitted
 12

 13       public boolean hasNext() {
 14         return (this.next == null);
 15       }
 16

 17   }

Code Sample 4.2: A linked list node Java implementation. Getter and setter methods
                 have been omitted for readability. A convenience method to determine
                 if a node has a next element is included. This implementation uses
                  null as its terminating value.




                                                                                  89
5 Algorithm Analysis

5.1 Introduction

An algorithm is a procedure or description of a procedure for solving a problem. An
algorithm is a step-by-step specification of operations to be performed in order to compute
an output, process data, or perform a function. An algorithm must always be correct (it
must always produce a valid output) and it must be finite (it must terminate after a
finite number of steps).
Algorithms are not code. Programs and code in a particular language are implementations
of algorithms. The word, “algorithm” itself is derived from the latinization of Abū
‘Abdalāh Muhammad ibn Mūsā al-Khwārizmı̄, a Persian mathematician (c. 780 – 850).
The concept of algorithms predates modern computers by several thousands of years.
Euclid’s algorithm for computing the greatest common denominator (see Section 5.5.3)
is 2,300 years old.
Often, to be useful an algorithm must also be feasible: given its input, it must execute
in a reasonable amount of time using a reasonable amount of resources. Depending on
the application requirements our tolerance may be on the order of a few milliseconds
to several days. An algorithm that takes years or centuries to execute is certainly not
considered feasible.

Deterministic An algorithm is deterministic if, when given a particular input, will always
     go through the exact same computational process and produce the same output.
     Most of the algorithms you’ve used up to this point are deterministic.
Randomized An algorithm that is randomized is an algorithm that involves some form
    of random input. The random source can be used to make decisions such as random
    selections or to generate random state in a program as candidate solutions. There
    are many types of randomized algorithms including Monte-Carlo algorithms (that
    may have some error with low probability), Las Vagas algorithms (whose results
    are always correct, but may fail with a certain probability to produce any results),
    etc.
Optimization Many algorithms seek not only to find a solution to a problem, but to
    find the best, optimal solution. Many of these type of algorithms are heuristics:
    rather than finding the actual best solution (which may be infeasible), they can
    approximate a solution (Approximation algorithms). Other algorithms simulate



                                                                                        91
5 Algorithm Analysis

     biological processes (Genetic algorithms, Ant Colony algorithms, etc.) to search
     for an optimal solution.
Parallel Most modern processors are multicore, meaning that they have more than one
      processor on a chip. Many servers have dozens of processors that work together.
      Multiple processors can be utilized by designing parallel algorithms that can split
     work across multiple processes or threads which can be executed in parallel to each
      other, improving overall performance.
Distributed Computation can also be distributed among completely separate devices that
      may be located half way across the globe. Massive distributed computation networks
      have been built for research such as simulating protein folding (Folding@Home).

An algorithm is a more abstract, generalization of what you might be used to in a
typical programming language. In an actual program, you may have functions/methods,
subroutines or procedures, etc. Each one of these pieces of code could be considered an
algorithm in and of itself. The combination of these smaller pieces create more complex
algorithms, etc. A program is essentially a concrete implementation of a more general,
theoretical algorithm.
When a program executes, it expends some amount of resources. For example:

Time The most obvious resource an algorithm takes is time: how long the algorithm
    takes to finish its computation (measured in seconds, minutes, etc.). Alternatively,
    time can be measured in how many CPU cycles or floating-point operations a
    particular piece of hardware takes to execute the algorithm.
Memory The second major resource in a computer is memory. An algorithm requires
   memory to store the input, output, and possibly extra memory during its execution.
   How much memory an algorithm uses in its execution may be even more of
   an important consideration than time in certain environments or systems where
   memory is extremely limited such as embedded systems.
Power The amount of power a device consumes is an important consideration when you
    have limited capacity such as a battery in a mobile device. From a consumer’s
    perspective, a slower phone that offered twice the battery life may be preferable.
    In certain applications such as wireless sensor networks or autonomous systems
    power may be more of a concern than either time or memory.
Bandwidth In computer networks, efficiency is measured by how much data you can
    transmit from one computer to another, called throughput. Throughput is generally
    limited by a network’s bandwidth: how much a network connection can transmit
    under ideal circumstances (no data loss, no retransmission, etc.)
Circuitry When designing hardware, resources are typically measured in the number
     of gates or wires are required to implement the hardware. Fewer gates and wires
     means you can fit more chips on a silicon die which results in cheaper hardware.
     Fewer wires and gates also means faster processing.



92
                                                                          5.1 Introduction

Idleness Even when a computer isn’t computing anything, it can still be “costing” you
     something. Consider purchasing hardware that runs a web server for a small user
     base. There is a substantial investment in the hardware which requires maintenance
     and eventually must be replaced. However, since the user base is small, most of
     the time it sits idle, consuming power. A better solution may be to use the same
     hardware to serve multiple virtual machines (VMs). Now several small web serves
     can be served with the same hardware, increasing our utilization of the hardware.
     In scenarios like this, the lack of work being performed is the resource.
Load Somewhat the opposite of idleness, sometimes an application or service may have
    occasional periods of high demand. The ability of a system to service such high
    loads may be considered a resource, even if the capacity to handle them goes unused
    most of the time.

These are all very important engineering and business considerations when designing
systems, code, and algorithms. However, we’ll want to consider the complexity of
algorithms in a more abstract manner.
Suppose we have two different programs (or algorithms) A and B. Both of those algorithms
are correct, but A uses fewer of the above resources than B. Clearly, algorithm A is the
better, more efficient solution. However, how can we better quantify this efficiency?


List Operations

To give a concrete example, consider the list data structures from Chapter 4. The list
could be implemented as an array-based list (where the class owns a static array that is
resized/copied when full) or a linked list (with nodes containing elements and linking to
the next node in the list). Some operations are “cheap” on one type of list while other
operations may be more “expensive.”
Consider the problem of inserting a new element into the list at the beginning (at index
0). For a linked list this involves creating a new node and shuffling a couple of references.
The number of operations in this case is not contingent on the size of the the list. In
contrast, for an array-based list, if the list contains n elements, each element will need to
be shifted over one position in the array in order to make room for the element to be
inserted. The number of shifts is proportional to the number of elements in the array, n.
Clearly for this operation, a linked list is better (more efficient).
Now consider a different operation: given an index i, retrieve the i-th element in the
list. For an array-based list we have the advantage of having random access to the array.
When we index an element, arr[i] , it only takes one memory address computation
to “jump” to the memory location containing the i-th element. In contrast, a linked list
would require us to start at the head, and traverse the list until we reach the i-th node.
This requires i traversal operations. In the worst case, retrieving the last element, the
n-th element, would require n such operations. A summary of these operations can be



                                                                                          93
5 Algorithm Analysis

                List Type          Insert at start   Index-based Retrieve
                Array-based List   n                 1
                Linked List        2                 i≈n

               Table 5.1: Summary of the Complexity of List Operations

  1   public static int sum(List<Integer> items) {
  2

  3       int total = 0;
  4       for(int i=0; i<items.size(); i++) {
  5         total += items.get(i);
  6       }
  7       return total;
  8   }

                  Code Sample 5.1: Summing a collection of integers

found in Table 5.1.
We will now demonstrate the consequences of this difference in performance for an
index-based retrieval operation. Consider the Java code in Code Sample 5.1. This
method is using a naive index-based retrieval (line 5) to access each element.
Suppose that the List passed to this method is an ArrayList in which each get(i)
method call take a constant number of operations and thus roughly a constant amount
of time. Since we perform this operation once for each element in the list, we can assume
that the amount of time that the entire algorithm will take will be proportional to n,
the number of elements in the list. That is, the time t that the algorithm will take to
execute will be some linear function of n, say

                                    tlin (n) = an + b

Here, the constants a and b are placeholders for some values that are dependent on the
particular machine that we run it on. These two values may vary depending on the
speed of the machine, the available memory, etc. and will necessarily change if we run
the algorithm on a different machine with different specifications. However, the overall
complexity of the algorithm, the fact that it takes a linear amount of work to execute,
will not change.
Now let’s contrast the scenario where a LinkedList is passed to this method instead.
On each iteration, each get(i) method will take i node traversals to retrieve the i-th
element. On the first iteration, only a single node traversal will be required. On the
second, 2 traversals are required, etc. Adding all of these operations together gives us
the following
                            1 + 2 + 3 + 4 + · · · + (n − 1) + n



94
                                                                                   5.1 Introduction

That is, the sum of natural numbers 1 up to n. We can use the well-known Gauss’s
Formula to get a closed form of this summation.
Theorem 1 (Gauss’s Formula).
                                            n
                                            X        n2 + n
                                                  i=
                                            i=1
                                                        2

Thus, the total number of operations will be proportional to some quadratic function,
                                       tquad (n) + an2 + bn + c
Again, we don’t know what the constants, a, b, c are as they may vary depending on the
languages, system and other factors. However, we can compute them experimentally.
The code in Code Sample was run on a laptop and its performance was timed. The
experiment generated random lists of various sizes, starting at n = 50,000 up to 1 million
by increments of 50,000. The time to execute was recorded in seconds. The experiment
was repeated for both types of lists multiple times and an average was taken to reduce
the influence of other factors (such as other processes running on the machine at the
same time).
To find the coefficients in both functions, a linear and quadratic regression was performed,
giving us the following functions
                                tlin (n) = 5.138e−5n + 0.004
                              tquad (n) = 6.410e−4n2 − 0.112n + 9.782
Both had very high correlation coefficients (the quadratic regression was nearly perfect at
0.994) which gives strong empirical evidence for our theoretical analysis. The quadratic
regression along with the experimental data points is given in Figure 5.1. The linear
data is not graphed as it would remain essentially flat on this scale.
These experimental results allow us to estimate and project how this code will perform
on larger and larger lists. For example, we can plug n = 10 million into tlin (n) and find
that it will still take less than 1 second for the method to execute. Likewise, we can find
out how long it would take for the linked list scenario. For a list of the same size, n = 10
million, the algorithm would take 17.498 hours! More values are depicted in Table 5.2.
The contrast between the two algorithms gets extreme even with moderately large lists of
integers. With 10 billion numbers, the quadratic approach is not even feasible. Arguably,
even at 10 million, a run time of over 17 hours is not acceptable, especially when an
alternative algorithm can perform the same calculation in less than 1 second.
Being able to identify the complexity of algorithms and the data structures they use in
order to avoid such inefficient solutions is an essential skill in Computers Science.1 In the
following examples, we’ll begin to be a little bit more formal about this type of analysis.
 1
     Using an index-based iteration is a common mistake in Java code. The proper solution would have
      been to use an enhanced for-loop or iterator pattern so that the method would perform the same on
      either array-based or linked lists.




                                                                                                    95
5 Algorithm Analysis




                                                    ���������������������������������
                     ����
                                   ����������
                                     �������
                     ����


                     ����
        ����������




                     ����


                     ����


                     ����


                         ��
                              ��                ����         ����                   ����   �   ���   �   ����
                                                                    �������������


     Figure 5.1: Quadratic Regression of Index-Based Linked List Performance.




                         Table 5.2: Various Projected Runtimes for Code Sample 5.1.
                                               Execution Time
                                  List Size
                                               Linear        Quadratic
                                  10 million   1 second      17.498 hours
                                  100 million 10 seconds     74.06 days
                                  1 billion    2 minutes     20.32 years
                                  10 billion   8.56 minutes 2031.20 years




96
                                                                       5.1 Introduction


  1   int result = 0;
  2   for(int i=1; i<=n; i++) {
  3     for(int j=1; j<=i; j++) {
  4       result = result + 1;
  5     }
  6   }

                      Code Sample 5.2: Summation Algorithm 1

  1   int result = 0;
  2   for(int i=1; i<=n; i++) {
  3     result = result + i;
  4   }

                      Code Sample 5.3: Summation Algorithm 2


5.1.1 Example: Computing a Sum

The following is a toy example, but its easy to understand and straightforward. Consider
the following problem: given an integer n ≥ 0, we want to compute the arithmetic series,
                         n
                         X
                                i = 1 + 2 + 3 + · · · + (n − 1) + n
                          i=1

As a naive approach, consider the algorithm in Code Sample 5.2. In this algorithm, we
iterate over each possible number i in the series. For each number i, we count 1 through
i and add one to a result variable.
As an improvement, consider the algorithm in Code Sample 5.3. Instead of just adding
one on each iteration of the inner loop, we omit the loop entirely and simply just add
the index variable i to the result.
Can we do even better? Yes. Recall Theorem 1 that gives us a direct closed form solution
for this summation:
                                    n
                                   X      n(n + 1)
                                       i=
                                   i=1
                                             2
Code Sample 5.4 uses this formula to directly compute the sum without any loops.


  1   int result = n * (n + 1) / 2;

                      Code Sample 5.4: Summation Algorithm 3



                                                                                     97
5 Algorithm Analysis

  Algorithm   Number of     Input Size
              Additions
                            10           100       1,000       10,000     100,000     1,000,000
  1           ≈ n2          0.003ms      0.088ms   1.562ms   2.097ms    102.846ms   9466.489ms
  2           n             0.002ms      0.003ms   0.020ms   0.213ms      0.872ms      1.120ms
  3           1             0.002ms      0.001ms   0.001ms   0.001ms      0.001ms      0.000ms

        Table 5.3: Empirical Performance of the Three Summation Algorithms


All three of these algorithms were run on a laptop computer for various values of n from
10 up to 1,000,000. Table 5.3 contains the resulting run times (in milliseconds) for each
of these three algorithms on the various input sizes.
With small input sizes, there is almost no difference between the three algorithms.
However, that would be a naive way of analyzing them. We are more interested in how
each algorithm performs as the input size, n increases. In this case, as n gets larger, the
differences become very stark. The first algorithm has two nested for loops. On average,
the inner loop will run about n2 times while the outer loop runs n times. Since the loops
are nested, the inner loop executes about n2 times for each iteration of the outer loop.
Thus, the total number of iterations, and consequently the total number of additions is
about
                                            n
                                       n × ≈ n2
                                            2
The second algorithm saves the inner for loop and thus only makes n additions. The
final algorithm only performs a constant number of operations.
Observe how the running time grows as the input size grows. For Algorithm 1, increasing
n from 100,000 to 1,000,000 (10 times as large) results in a running time that is about
100 times as slow. This is because it is performing n2 operations. To see this, consider
the following. Let t(n) be the time that Algorithm 1 takes for an input size of n. From
before we know that
                                        t(n) ≈ n2
Observe what happens when we increase the input size from n to 10n:

                                 t(10n) ≈ (10n)2 = 100n2

which is 100 times as large as t(n). The running time of Algorithm 1 will grow quadratically
with respect to the input size n.
Similarly, Algorithm 2 grows linearly,

                                            t(n) ≈ n

Thus, a 10 fold increase in the input,

                                          t(10n) ≈ 10n



98
                                                                                       5.1 Introduction


     1   public static int mode01(int arr[]) {
     2

     3       int maxCount = 0;
     4       int modeIndex = 0;
     5       for(int i=0; i<arr.length; i++) {
     6         int count = 0;
     7         int candidate = arr[i];
     8         for(int j=0; j<arr.length; j++) {
     9           if(arr[j] == candidate) {
 10                count++;
 11              }
 12            }
 13            if(count > maxCount) {
 14              modeIndex = i;
 15              maxCount = count;
 16            }
 17          }
 18          return arr[modeIndex];
 19      }

                           Code Sample 5.5: Mode Finding Algorithm 1


leads to a 10 fold increase in the running time. Algorithm 3’s runtime does not depend
on the input size, and so its runtime does not grow as the input size grows. It essentially
remains flat–constant.
Of course, the numbers in Table 5.3 don’t follow this trend exactly, but they are pretty
close. The actual experiment involves a lot more variables than just the algorithms: the
laptop may have been performing other operations, the compiler and language may have
optimizations that change the algorithms, etc. Empirical results only provide general
evidence as to the runtime of an algorithm. If we moved the code to a different, faster
machine or used a different language, etc. we would get different numbers. However, the
general trends in the rate of growth would hold. Those rates of growth will be what we
want to analyze.



5.1.2 Example: Computing a Mode

As another example, consider the problem of computing the mode of a collection of
numbers. The mode is the most common element in a set of data.2
 2
     In general there may be more than one mode, for example in the set {10, 20, 10, 20, 50}, 10 and 20 are
      both modes. The problem will simply focus on finding a mode, not all modes.




                                                                                                        99
5 Algorithm Analysis


  1   public static int mode02(int arr[]) {
  2     Arrays.sort(arr);
  3     int i=0;
  4     int modeIndex = 0;
  5     int maxCount = 0;
  6     while(i < arr.length-1) {
  7       int count=0;
  8       while(i < arr.length-1 && arr[i] == arr[i+1]) {
  9         count++;
 10         i++;
 11       }
 12       if(count > maxCount) {
 13         modeIndex = i;
 14         maxCount = count;
 15       }
 16       i++;
 17     }
 18     return arr[modeIndex];
 19   }

                      Code Sample 5.6: Mode Finding Algorithm 2

Consider the strategy as illustrated in Code Sample 5.5. For each element in the array,
we iterate through all the other elements and count how many times it appears (its
multiplicity). If we find a number that appears more times than the candidate mode
we’ve found so far, we update our variables and continue. As with the previous algorithm,
the nested nature of our loops leads to an algorithm that performs about n2 operations
(in this case, the comparison on line 9).
Now consider the following variation in Code Sample 5.6. In this algorithm, the first
thing we do is sort the array. This means that all equal elements will be contiguous. We
can exploit this to do less work. Rather than going through the list a second time for
each possible mode, we can count up contiguous runs of the same element. This means
that we need only examine each element exactly once, giving us n comparison operations
(line 8).
We can’t, however, ignore the fact that to exploit the ordering, we needed to first “invest”
some work upfront by sorting the array. Using a typical sorting algorithm, we would
expect that it would take about n log (n) comparisons. Since the sorting phase and mode
finding phase were separate, the total number of comparisons is about
                                       n log (n) + n
The highest order term here is the n log (n) term for sorting. However, this is still lower
than the n2 algorithm. In this case, the investment to sort the array pays off! To compare



100
                                                                         5.1 Introduction


  1    public static int mode03(int arr[]) {
  2      Map<Integer, Integer> counts = new HashMap<Integer, Integer>();
  3      for(int i=0; i<arr.length; i++) {
  4        Integer count = counts.get(arr[i]);
  5        if(count == null) {
  6          count = 0;
  7        }
  8        count++;
  9        counts.put(arr[i], count);
 10      }
 11      int maxCount = 0;
 12      int mode = 0;
 13      for(Entry<Integer, Integer> e : counts.entrySet()) {
 14        if(e.getValue() > maxCount) {
 15          maxCount = e.getValue();
 16          mode = e.getKey();
 17        }
 18      }
 19      return mode;
 20    }

                     Code Sample 5.7: Mode Finding Algorithm 3


with our previous analysis, what happens when we increase the input size 10 fold? For
simplicity, let’s only consider the highest order term:

                                     t(n) = n log (n)

Then
                   t(10n) = 10n log (10n) = 10n log (n) + 10n log (10)
Ignoring the lower order term, the increase in running time is essentially linear! We
cannot discount the additive term in general, but it is so close to linear that terms like
n log (n) are sometimes referred to as quasilinear.
Yet another solution, presented in Code Sample 5.7, utilizes a map data structure to
compute the mode. A map is a data structure that allows you to store key-value pairs.
In this case, we map elements in the array to a counter that represents the element’s
multiplicity. The algorithm works by iterating over the array and entering/updating the
elements and counters.
There is some cost associated with inserting and retrieving elements from the map,
but this particular implementation offers amortized constant running time for these
operations. That is, some particular entries/retrievals may be more expensive (say



                                                                                      101
5 Algorithm Analysis

 Algorithm   Number of    Input Size
             Additions
                          10           100       1,000         10,000      100,000       1,000,000
 1           ≈ n2         0.007ms      0.155ms   11.982ms   45.619ms    3565.570ms   468086.566ms
 2           n            0.143ms      0.521ms   2.304ms    19.588ms      40.038ms      735.351ms
 3           n            0.040ms      0.135ms   0.703ms    10.386ms      21.593ms      121.273ms

        Table 5.4: Empirical Performance of the Three Mode Finding Algorithms


linear), but when averaged over the life of the algorithm/data structure, each operation
only takes a constant amount of time.
Once built, we need only go through the elements in the map (at most n) and find the
one with the largest counter. This algorithm, too, offers essentially linear runtime for all
inputs. Similar experimental results can be found in Table 5.4.
The difference in performance is even more dramatic than in the previous example. For an
input size of 1,000,000 elements, the n2 algorithm took nearly 8 minutes! This is certainly
unacceptable performance for most applications. If we were to extend the experiment to
n = 10,000,000, we would expect the running time to increase to about 13 hours! For
perspective, input sizes in the millions are small by today’s standards. Algorithms whose
runtime is quadratic are not considered feasible for today’s applications.



5.2 Pseudocode

We will want to analyze algorithms in an abstract, general way independent of any
particular hardware, framework, or programming language. In order to do this, we need
a way to specify algorithms that is also independent of any particular language. For that
purpose, we will use pseudocode.
Pseudocode (“fake” code) is similar to some programming languages that you’re familiar
with, but does not have any particular syntax rules. Instead, it is a higher-level description
of a process. You may use familiar control structures such as loops and conditionals, but
you can also utilize natural language descriptions of operations.
There are no established rules for pseudocode, but in general, good pseudocode:
     • Clearly labels the algorithm
     • Identifies the input and output at the top of the algorithm
     • Does not involve any language or framework-specific syntax–no semicolons, decla-
       ration of variables or their types, etc.
     • Makes liberal use of mathematical notation and natural language for clarity
Good pseudocode abstracts the algorithm by giving enough details necessary to under-



102
                                                                                      5.2 Pseudocode

stand the algorithm and subsequently implement it in an actual programming language.
Let’s look at some examples.


         Input : A collection of numbers, A = {a1 , . . . , an }
         Output : The mean, µ of the values in A
     1   sum ← 0
     2   foreach ai ∈ A do
     3      sum ← sum + ai
     4   end
     5   µ ← sum
               n
     6   output µ

                            Algorithm 5: Computing the Mean

Algorithm 5 describes a way to compute the average of a collection of numbers. Observe:
      • The input does not have a specific type (such as int or double ), it uses set
        notation which also indicates how large the collection is.
      • There is no language-specific syntax such as semicolons, variable declarations, etc.
      • The loop construct doesn’t specify the details of incrementing a variable, instead
        using a “foreach” statement with some set notation3
      • Code blocks are not denoted by curly brackets, but are clearly delineated by using
        indentation and vertical lines.
      • Assignment and compound assignment operators do not use the usual syntax from
        C-style languages, instead using a left-oriented arrow to indicate a value is assigned
        to a variable.4
Consider another example of computing the mode, similar to the second approach in a
previous example.
Some more observations about Algorithm 6:
      • The use of natural language to specify that the collection should be sorted and in
        what order
      • The usage of −∞ as a placeholder so that any other value would be greater than it
      • The use of natural language to specify that an iteration takes place over contiguous
        elements (line 3) or that a sub-operation such as a count/summation (line 4) is
        performed
 3
     To review, ai ∈ A is a predicate meaning the element ai is in the set A.
 4
     Not all languages use the familiar single equals sign = for the assignment operator. The statistical
      programming language R uses the left-arrow operator, <- and Maple uses := for example.




                                                                                                    103
5 Algorithm Analysis


      Input : A collection of numbers, A = {a1 , . . . , an }
      Output : A mode of A
  1   Sort the elements in A in non-decreasing order
  2   multiplicity ← −∞
  3   foreach run of contiguous equal elements a do
  4      m ← count up the number of times a appears
  5      if m > multiplicity then
  6          mode ← a
  7          multiplicity ← m
  8      end
  9   end
 10   output m

                        Algorithm 6: Computing the Mode

In contrast, bad pseudocode would have the opposite elements. Writing a full program
or code snippet in Java for example. Bad pseudocode may be unclear or it may overly
simplify the process to the point that the description is trivial. For example, suppose we
wanted to specify a sorting algorithm, and we did so using the pseudocode in Algorithm
7. This trivializes the process. There are many possible sorting algorithms (insertion
sort, quick sort, etc.) but this algorithm doesn’t specify any details for how to go about
sorting it.
On the other hand, in Algorithm 6, we did essentially do this. In that case it was perfectly
fine: sorting was a side operation that could be achieved by a separate algorithm. The
point of the algorithm was not to specify how to sort, but instead how sorting could be
used to solve another problem, finding the mode.

      Input : A collection of numbers, A = {a1 , . . . , an }
      Output : A0 , sorted in non-decreasing order
  1   A0 ← Sort the elements in A in non-decreasing order
  2   output A0

                 Algorithm 7: Trivial Sorting (Bad Pseudocode)

Another example would be if we need to find a minimal element in a collection. Trivial
pseudocode may be like that found in Algorithm 8. No details are presented on how to
find the element. However, if finding the minimal element were an operation used in
a larger algorithm (such as selection sort), then this terseness is perfectly fine. If the
primary purpose of the algorithm is to find the minimal element, then details must be
presented as in Algorithm 9.



104
                                                                           5.3 Analysis


      Input : A collection of numbers, A = {a1 , . . . , an }
      Output : The minimal element of A
  1   m ← minimal element of A
  2   output m

              Algorithm 8: Trivially Finding the Minimal Element



      Input : A collection of numbers, A = {a1 , . . . , an }
      Output : The minimal element of A
  1   m←∞
  2   foreach ai ∈ A do
  3      if ai < m then
  4          m ← ai
  5      end
  6   end
  7   output m

                   Algorithm 9: Finding the Minimal Element



5.3 Analysis

Given two competing algorithms, we could empirically analyze them like we did in
previous examples. However, it may be infeasible to implement both just to determine
which is better. Moreover, by analyzing them from a more abstract, theoretical approach,
we have a better more mathematically-based proof of the relative complexity of two
algorithm.
Given an algorithm, we can analyze it by following this step-by-step process.
  1. Identify the input
  2. Identify the input size, n
  3. Identify the elementary operation
  4. Analyze how many times the elementary operation is executed with respect to the
     input size n
  5. Characterize the algorithm’s complexity by providing an asymptotic (Big-O, or
     Theta) analysis



                                                                                    105
5 Algorithm Analysis

Identifying the Input

This step is pretty straightforward. If the algorithm is described with good pseudocode,
then the input will already be identified. Common types of inputs are single numbers,
collections of elements (lists, arrays, sets, etc.), data structures such as graphs, matrices,
etc.
However, there may be some algorithms that have multiple inputs: two numbers or a
collection and a key, etc. In such cases, it simplifies the process if you can, without loss
of generality, restrict attention to a single input value, usually the one that has the most
relevance to the elementary operation you choose.


Identifying the Input Size

Once the input has been identified, we need to identify its size. We’ll eventually want to
characterize the algorithm as a function f (n): given an input size, how many resources
does it take. Thus, it is important to identify the number corresponding to the domain
of this function.
This step is also pretty straightforward, but may be dependent on the type of input or
even its representation. Examples:
   • For collections (sets, lists, arrays), the most natural is to use the number of elements
     in the collection (cardinality, size, etc.). The size of individual elements is not as
     important as number of elements since the size of the collection is likely to grow
     more than individual elements do.
   • An n × m matrix input could be measured by one or both nm of its dimensions.
   • For graphs, you could count either the number of vertices or the number of edges
     in the graph (or both!). How the graph is represented may also affect its input size
     (an adjacency matrix vs. an adjacency list).
   • If the input is a number x, the input size is typically the number of bits required
     to represent x. That is,
                                        n ≈ log2 (x)
      To see why, recall that if you have n bits, the maximum unsigned integer you can
      represent is 2n − 1. Inverting this expression gives us dlog2 (x + 1)e.
Some algorithms may have multiple inputs. For example, a collection and a number (for
searching) or two integers as in Euclid’s algorithm. The general approach to analyzing
such algorithms is to simplify things by only considering one input. If one of the inputs
is larger, such as a collection vs. a single element, the larger one is used in the analysis.
Even if it is not clear which one is larger, it may be possible to assume, without loss of
generality, that one is larger than the other (and if not, the inputs may be switched).
The input size can then be limited to one variable to simplify the analysis.



106
                                                                              5.3 Analysis

Identifying the Elementary Operation

We also need to identify what part of the algorithm does the actual work (where the most
resources will be expended). Again, we want to keep the analysis simple, so we generally
only identify one elementary operation. There may be several reasonable candidates
for the elementary operation, but in general it should be the most common or most
expensive operation performed in the algorithm. For example:
   • When performing numeric computations, arithmetic operations such as additions,
     divisions, etc.
   • When sorting or searching, comparisons are the most natural elementary operations.
     Swaps may also be a reasonable choice depending on how you want to analyze the
     algorithm.
   • When traversing a data structure such as a linked list, tree, or graph a node traversal
     (visiting or processing a node) may be considered the elementary operation.
In general, operations that are necessary to control structures (such as loops, assignment
operators, etc.) are not considered good candidates for the elementary operation. An
extended discussion of this can be found in Section 5.6.2.


Analysis

Once the elementary operation has been identified, the algorithm must be analyzed to
count the number of times it is executed with respect to the input size. That is, we
analyze the algorithm to find a function f (n) where n is the input size and f (n) gives
the number of times the elementary operation is executed.
The analysis may involve deriving and solving a summation. For example, if the
elementary operation is performed within a for loop and the loop runs a number of times
that depends on the input size n.
If there are multiple loops in which the elementary operation is performed, it may be
necessary to setup multiple summations. If two loops are separate and independent (one
executes after the other), then the sum rule applies. The total number of operations is
the sum of the operations of each loop.
If two loops are nested, then the product rule applies. The inner loop will execute fully
for each iteration of the outer loop. Thus, the number of operations are multiplied with
each other.
Sometimes the analysis will not be so clear cut. For example, a while loop may execute
until some condition is satisfied that does not directly depend on the input size but also
on the nature of the input. In such cases, we can simplify our analysis by considering
the worst-case scenario. In the while loop, what is the maximum possible number of
iterations for any input?



                                                                                        107
5 Algorithm Analysis

                             ·106
                       1.5



                        1
                                    f (n) = 100n2 + 5n


                       0.5

                                                          g(n) = n3
                        0
                             0      20     40    60      80   100     120
                                                 n

                                 Figure 5.2: Plot of two functions.


Asymptotic Characterization

As computers get faster and faster and resources become cheaper, they can process more
and more information in the same amount of time. However, the characterization of
an algorithm should be invariant with respect to the underlying hardware. If we run
an algorithm on a machine that is twice as fast, that doesn’t mean that the algorithm
has improved. It still takes the same number of operations to execute. Faster hardware
simply means that the time it takes to execute those operations is half as much as it was
before.
To put it in another perspective, performing Euclid’s algorithm to find the GCD of two
integers took the same number of steps 2,300 years ago when he performed them on
paper as it does today when they are executed on a digital computer. A computer is
obviously faster than Euclid would have been, but both Euclid and the computer are
performing the same number of steps when executing the same algorithm.
 For this reason, we characterize the number of operations performed by an algorithm
 using asymptotic analysis. Improving the hardware by a factor of two only affects the
“hidden constant” sitting outside of the function produced by the analysis in the previous
 step. We want our characterization to be invariant of those constants.
Moreover, we are really more interested in how our algorithm performs for larger and
larger input sizes. To illustrate, suppose that we have two algorithms, one that performs

                                         f (n) = 100n2 + 5n

operations and one that performs
                                             g(n) = n3



108
                                                                                     5.4 Asymptotics

operations. These functions are graphed in Figure 5.2. For inputs of size less than 100,
the first algorithm performs worse than the second (the graph is higher indicating “more”
resources). However, for inputs of size greater than 100, the first algorithm is better. For
small inputs, the second algorithm may be better, but small inputs are not the norm
for any “real” problems.5 In any case, on modern computers, we would expect small
inputs to execute fast anyway as they did in our empirical experiments in Section 5.1.1
and 5.1.2. There was essentially no discernible difference in the three algorithms for
sufficiently small inputs.
We can rigorously quantify this by providing an asymptotic characterization of these
functions. An asymptotic characterization essentially characterizes the rate of growth of
a function or the relative rate of growth of functions. In this case, n3 grows much faster
than 100n2 + 5n as n grows (tends toward infinity). We formally define these concepts
in the next section.



5.4 Asymptotics

5.4.1 Big-O Analysis

We want to capture the notion that one function grows faster than (or at least as fast as)
another. Categorizing functions according to their growth rate has been done for a long
time in mathematics using big-O notation.6

Definition 1. Let f and g be two functions, f, g : N → R+ . We say that

                                           f (n) ∈ O(g(n))

read as “f is big-O of g,” if there exist constants c ∈ R+ and n0 ∈ N such that for every
integer n ≥ n0 ,
                                        f (n) ≤ cg(n)

First, let’s make some observations about this definition.
     • The “O” originally stood for “order of”, Donald Knuth referred to it as the capital
       greek letter omicron, but since it is indistinguishable from the Latin letter “O” it
       makes little difference.
 5
   There are problems where we can apply a “hybrid” approach: we can check for the input size and
    choose one algorithm for small inputs and another for larger inputs. This is typically done in hybrid
    sorting algorithms such as when merge sort is performed for “large” inputs but switches over to
    insertion sort for smaller arrays.
 6
   The original notation and definition are attributed to Paul Bachmann in 1894 [2]. Definitions and
    notation have been refined and introduced/reintroduced over the years. Their use in algorithm
    analysis was first suggested by Donald Knuth in 1976 [7].




                                                                                                    109
5 Algorithm Analysis

   • Some definitions are more general about the nature of the functions f, g. However,
     since we’re looking at these functions as characterizing the resources that an
     algorithm takes to execute, we’ve restricted the domain and codomain of the
     functions. The domain is restricted to non-negative integers since there is little
     sense in negative or factional input sizes. The codomain is restricted to nonnegative
     reals as it doesn’t make sense that an algorithm would potentially consume a
     negative amount of resources.
   • We’ve used the set notation f (n) ∈ O(g(n)) because, strictly speaking, O(g(n)) is
     a class of functions: the set of all functions that are asymptotically bounded by
     g(n). Thus the set notation is the most appropriate. However, you will find many
     sources and papers using notation similar to

                                         f (n) = O(g(n))

      This is a slight abuse of notation, but common nonetheless.
The intuition behind the definition of big-O is that f is asymptotically less than or equal
to g. That is, the rate of growth of g is at least as fast as the growth rate of f . Big-O
provides a means to express that one function is an asymptotic upper bound to another
function.
The definition essentially states that f (n) ∈ O(g(n)) if, after some point (for all n ≥ n0 ),
the value of the function g(n) will always be larger than f (n). The constant c possibly
serves to “stretch” or “compress” the function, but has no effect on the growth rate of
the function.


Example

Let’s revisit the example from before where f (n) = 100n2 + 5n and g(n) = n3 . We want
to show that f (n) ∈ O(g(n)). By the definition, we need to show that there exists a c
and n0 such that
                                      f (n) ≤ cg(n)
As we observed in the graph in Figure 5.2, the functions “crossed over” somewhere
around n = 100. Let’s be more precise about that. The two functions cross over when
they are equal, so we setup an equality,

                                      100n2 + 5n = n3

Collecting terms and factoring out an n (that is, the functions have one crossover point
at n = 0), we have
                                  n2 − 100n − 5 = 0
The values of n satisfying this inequality can be found by applying the quadratic formula,
and so                                       √
                                      100 ± 10000 + 20
                                  n=
                                                2


110
                                                                         5.4 Asymptotics

Which is −0.049975 . . . and 100.0499 . . .. The first root is negative and so irrelevant.
The second is our cross over point. The next largest integer is 101. Thus, for c = 1 and
n0 = 101, the inequality is satisfied.
In this example, it was easy to find the intersection because we could employ the quadratic
equation to find roots. This is much more difficult with higher degree polynomials. Throw
in some logarithmic functions, exponential functions, etc. and this approach can be
difficult.
Revisit the definition of big-O: the inequality doesn’t have to be tight or precise. In the
previous example we essentially fixed c and tried to find n0 such that the inequality held.
Alternatively, we could fix n0 to be small and then find the c (essentially compressing
the function) such that the inequality holds. Observe:

            100n2 + 5n ≤ 100n2 + 5n2                since n ≤ n2 for all n ≥ 0
                       = 105n2
                       ≤ 105n3                     since n2 ≤ n3 for all n ≥ 0
                       = 105g(n)


By adding positive values, we make the equation larger until it looks like what we want,
in this case g(n) = n3 . By the end we’ve got our constants: for c = 105 and n0 = 0, the
inequality holds. There is nothing special about this c, c = 1000000 would work too.
The point is we need only find at least one c, n0 pair that the inequality holds (there are
an infinite number of possibilities).


5.4.2 Other Notations

Big-O provides an asymptotic upper bound characterization of two functions. There are
several other notations that provide similar characterizations.


Big-Omega

Definition 2. Let f and g be two functions, f, g : N → R+ . We say that

                                     f (n) ∈ Ω(g(n))

read as “f is big-Omega of g,” if there exist constants c ∈ R+ and n0 ∈ N such that for
every integer n ≥ n0 ,
                                      f (n) ≥ cg(n)

Big-Omega provides an asymptotic lower bound on a function. The only difference is the
inequality has been reversed. Intuitively f has a growth rate that is bounded below by g.



                                                                                       111
5 Algorithm Analysis

Big-Theta

Yet another characterization can be used to show that two functions have the same order
of growth.

Definition 3. Let f and g be two functions f, g : N → R+ . We say that

                                     f (n) ∈ Θ(g(n))

read as “f is Big-Theta of g,” if there exist constants c1 , c2 ∈ R+ and n0 ∈ N such that
for every integer n ≥ n0 ,
                                 c1 g(n) ≤ f (n) ≤ c2 g(n)

Big-Θ essentially provides an asymptotic equivalence between two functions. The function
f is bounded above and below by g. As such, both functions have the same rate of
growth.


Soft-O Notation

Logarithmic factors contribute very little to a function’s rate of growth especially com-
pared to larger order terms. For example, we called n log (n) quasi linear since it was
nearly linear. Soft-O notation allows us to simplify terms by removing logarithmic factors.

Definition 4. Let f, g be functions such that f (n) ∈ O(g(n) · logk (n)). Then we say
that f (n) is soft-O of g(n) and write

                                     f (n) ∈ Õ(g(n))

For example,
                                    n log (n) ∈ Õ(n)


Little Asymptotics

Related to big-O and big-Ω are their corresponding “little” asymptotic notations, little-o
and little-ω.

Definition 5. Let f and g be two functions f, g : N → R+ . We say that

                                     f (n) ∈ o(g(n))

read as “f is little-o of g,” if
                                          f (n)
                                      lim       =0
                                      n→∞ g(n)




112
                                                                             5.4 Asymptotics

The little-o is sometimes defined as for every  > 0 there exists a constant N such that

                                 |f (n)| ≤ |g(n)| ∀n ≥ N

but given the restriction that g(n) is positive, the two definitions are essentially equivalent.
Little-o is a much stronger characterization of the relation of two functions. If f (n) ∈
o(g(n)) then not only is g an asymptotic upper bound on f , but they are not asymptoti-
cally equivalent. Intuitively, this is similar to the difference between saying that a ≤ b
and a < b. The second is a stronger statement as it implies the first, but the first does
not imply the second. Analogous to this example, little-o provides a “strict” asymptotic
upper bound. The growth rate of g is strictly greater than the growth rate of f .
Similarly, a little-ω notation can be used to provide a strict lower bound characterization.
Definition 6. Let f and g be two functions f, g : N → R+ . We say that

                                       f (n) ∈ ω(g(n))

read as “f is little-omega of g,” if
                                           f (n)
                                        lim      =∞
                                       n→∞ g(n)




5.4.3 Observations

As you might have surmised, big-O and big-Ω are duals of each other, thus we have the
following.
Lemma 1. Let f, g be functions. Then

                           f (n) ∈ O(g(n)) ⇐⇒ g(n) ∈ Ω(f (n))

Because big-Θ provides an asymptotic equivalence, both functions are big-O and big-Θ
of each other.
Lemma 2. Let f, g be functions. Then

                f (n) ∈ Θ(g(n)) ⇐⇒ f (n) ∈ O(g(n)) and f (n) ∈ Ω(g(n))

Equivalently,

                f (n) ∈ Θ(g(n)) ⇐⇒ g(n) ∈ O(f (n)) and g(n) ∈ Ω(f (n))

With respect to the relationship between little-o and little-ω to big-O and big-Ω, as
previously mentioned, little asymptotics provide a stronger characterization of the growth
rate of functions. We have the following as a consequence.



                                                                                           113
5 Algorithm Analysis

Lemma 3. Let f, g be functions. Then
                             f (n) ∈ o(g(n)) ⇒ f (n) ∈ O(g(n))
and
                             f (n) ∈ ω(g(n)) ⇒ f (n) ∈ Ω(g(n))

Of course, the converses of these statements do not hold.


Common Identities

As a direct consequence of the definition, constant coefficients in a function can be
ignored.
Lemma 4. For any constant c,
                                     c · f (n) ∈ O(f (n))

In particular, for c = 1, we have that
                                       f (n) ∈ O(f (n))
and so any function is an upper bound on itself.
In addition, when considering the sum of two functions, f1 (n), f2 (n), it suffices to consider
the one with a larger rate of growth.
Lemma 5. Let f1 (n), f2 (n) be functions such that f1 (n) ∈ O(f2 (n)). Then
                                 f1 (n) + f2 (n) ∈ O(f2 (n))

In particular, when analyzing algorithms with independent operations (say, loops), we
only need to consider the operation with a higher complexity. For example, when we
presorted an array to compute the mode, the presort phase was O(n log (n)) and the
mode finding phase was O(n). Thus the total complexity was
                                n log (n) + n ∈ O(n log (n))

When dealing with a polynomial of degree k,
                       ck nk + ck−1 nk−1 + ck−2 nk−2 + · · · + c1 n + c0
The previous results can be combined to conclude the following lemma.
Lemma 6. Let p(n) be a polynomial of degree k,
                   p(n) = ck nk + ck−1 nk−1 + ck−2 nk−2 + · · · + c1 n + c0
then
                                         p(n) ∈ Θ(nk )



114
                                                                            5.4 Asymptotics

     Class Name           Asymptotic Characterization          Algorithm Examples
     Constant             O(1)                                 Evaluating a formula
     Logarithmic          O(log (n))                           Binary Search
     Polylogarithmic      O(logk (n))
     Linear               O(n)                                 Linear Search
     Quasilinear          O(n log (n))                         Mergesort
     Quadratic            O(n2 )                               Insertion Sort
     Cubic                O(n3 )
     Polynomial           O(nk ) for any k > 0
     Exponential          O(2n )                               Computing a powerset
     Super-Exponential    O(2f (n) ) for f (n) ∈ Ω(n)          Computing permutations
                          For example, n!

                  Table 5.5: Common Algorithmic Efficiency Classes


Logarithms

When working with logarithmic functions, it suffices to consider a single base. As
Computer Scientists, we always work in base-2 (binary). Thus when we write log (n), we
implicitly mean log2 (n) (base-2). It doesn’t really matter though because all logarithms
are the same to within a constant as a consequence of the change of base formula:

                                                 loga (n)
                                    logb (n) =
                                                 loga (b)

That means that for any valid bases a, b,

                                  logb (n) ∈ Θ(loga (n))

Another way of looking at it is that an algorithm’s complexity is the same regardless of
whether or not it is performed by hand in base-10 numbers or on a computer in binary.
Other logarithmic identities that you may find useful remembering include the following:

                                   log (nk ) = k log (n)

                             log (n1 n2 ) = log (n1 ) + log (n2 )


Classes of Functions

Table 5.5 summarizes some of the complexity functions that are common when doing
algorithm analysis. Note that these classes of functions form a hierarchy. For example,
linear and quasilinear functions are also O(nk ) and so are polynomial.



                                                                                        115
5 Algorithm Analysis

5.4.4 Limit Method


The method used in previous examples directly used the definition to find constants c, n0
that satisfied an inequality to show that one function was big-O of another. This can get
quite tedious when there are many terms involved. A much more elegant proof technique
borrows concepts from calculus.
Let f (n), g(n) be functions. Suppose we examine the limit, as n → ∞ of the ratio of
these two functions.



                                            f (n)
                                         lim
                                        n→∞ g(n)




One of three things could happen with this limit.
The limit could converge to 0. If this happens, then by Definition 5 we have that
f (n) ∈ o(g(n)) and so by Lemma 3 we know that f (n) ∈ O(g(n)). This makes sense: if
the limit converges to zero that means that g(n) is growing much faster than f (n) and
so f is big-O of g.
The limit could diverge to infinity. If this happens, then by Definition 6 we have that
f (n) ∈ ω(g(n)) and so again by Lemma 3 we have f (n) ∈ Ω(g(n)). This also makes
sense: if the limit diverges, f (n) is growing much faster than g(n) and so f (n) is big-Ω
of g.
Finally, the limit could converge to some positive constant (recall that both functions are
restricted to positive codomains). This means that both functions have essentially the
same order of growth. That is, f (n) ∈ Θ(g(n). As a consequence, we have the following
Theorem.




Theorem 2 (Limit Method). Let f (n) and g(n) be functions. Then if



                                  
                                    0     then f (n) ∈ O(g(n))
                           f (n) 
                       lim      =   c > 0 then f (n) ∈ Θ(g(n))
                      n→∞ g(n)
                                    ∞     then f (n) ∈ Ω(g(n))
                                  



116
                                                                         5.4 Asymptotics

Examples

Let’s reuse the example from before where f (n) = 100n2 + 5n and g(n) = n3 . Setting up
our limit,
                                 f (n)       100n2 + 5n
                             lim       = lim
                            n→∞ g(n)     n→∞     n3
                                             100n + 5
                                       = lim
                                         n→∞    n2
                                             100n         5
                                       = lim       +  lim
                                         n→∞ n2      n→∞ n2
                                             100
                                       = lim     +0
                                         n→∞ n
                                       =0


And so by Theorem 2, we conclude that

                                      f (n) ∈ O(g(n))

Consider the following example: let f (n) = log2 n and g(n) = log3 (n2 ). Setting up our
limit we have
                                      f (n)    log2 n
                                lim         =
                                n→∞   g(n)    log3 n2
                                              log n
                                            = 2 log2 n
                                                  2
                                               log2 3
                                             log2 3
                                           =
                                                2
                                           = .7924 . . . > 0

And so we conclude that log2 (n) ∈ Θ(log3 (n2 )).
As another example, let f (n) = log (n) and g(n) = n. Setting up the limit gives us
                                            log (n)
                                        lim
                                        n→∞    n
The rate of growth might seem obvious here, but we still need to be mathematically
rigorous. Both the denominator and numerator are monotone increasing functions. To
solve this problem, we can apply l’Hôpital’s Rule:
Theorem 3 (l’Hôpital’s Rule). Let f and g be functions. If the limit of the quotient fg(n)
                                                                                        (n)

exists, it is equal to the limit of the derivative of the denominator and the numerator.
That is,
                                        f (n)        f 0 (n)
                                    lim       = lim 0
                                   n→∞ g(n)     n→∞ g (n)




                                                                                      117
5 Algorithm Analysis

Applying this to our limit, the denominator drops out, but what about the numerator?
Recall that log (n) is the logarithm base-2. The derivative of the natural logarithm is well
known, ln0 (n) = n1 . We can use the change of base formula to transform log (n) = ln      (n)
                                                                                        ln (2)
and then take the derivative. That is,
                                                        1
                                        log0 (n) =
                                                     ln (2)n
Thus,
                                      log (n)       log0 (n)
                                  lim         = lim
                                  n→∞    n      n→∞    n0
                                                       1
                                              = lim
                                                n→∞ ln (2)n

                                              =0
Concluding that log (n) ∈ O(n).


Pitfalls

l’Hôpital’s Rule is not always the most appropriate tool to use. Consider the following
example: let f (n) = 2n and g(n) = 3n . Setting up our limit and applying l’Hôpital’s
Rule we have
                                         2n       (2n )0
                                     lim    = lim n 0
                                    n→∞ 3n    n→∞ (3 )

                                                  (ln 2)2n
                                            = lim
                                              n→∞ (ln 3)3n




which doesn’t get us anywhere. In general, we should look for algebraic simplifications
first. Doing so we would have realized that
                                                 n
                                      2n         2
                                  lim n = lim
                                 n→∞ 3      n→∞  3
        2
Since   3
            < 1, the limit of its exponent converges to zero and we have that 2n ∈ O(3n ).


5.5 Examples

5.5.1 Linear Search

As a simple example, consider the problem of searching a collection for a particular
element. The straightforward solution is known as Linear Search and is featured as



118
                                                                             5.5 Examples

Algorithm 10

      Input : A collection A = {a1 , . . . , an }, a key k
      Output : The first i such that ai = k, φ otherwise
  1   for i = 1, . . . , n do
  2      if ai = k then
  3          output i
  4      end
  5   end
  6   output φ

                          Algorithm 10: Linear Search
Let’s follow the prescribed outline above to analyze Linear Search.
  1. Input: this is clearly indicated in the pseudocode of the algorithm. The input is
     the collection A.
  2. Input Size: the most natural measure of the size of a collection is its cardinality; in
     this case, n
  3. Elementary Operation: the most common operation is the comparison in line 2
     (assignments and iterations necessary for the control flow of the algorithm are not
     good candidates).
  4. How many times is the elementary operation executed with respect to the input
     size, n? The situation here actually depends not only on n but also the contents of
     the array.
         • Suppose we get lucky and find k in the first element, a1 : we’ve only made one
           comparison.
         • Suppose we are unlucky and find it as the last element (or don’t find it at all).
           In this case we’ve made n comparisons
         • We could also look at the average number of comparisons (see Section 5.6.3)
       In general, algorithm analysis considers at the worst case scenario unless otherwise
       stated. In this case, there are C(n) = n comparisons.
  5. This is clearly linear, which is why the algorithm is called Linear Search,




                                              Θ(n)



                                                                                        119
5 Algorithm Analysis

5.5.2 Set Operation: Symmetric Difference

Recall that the symmetric difference of two sets, A ⊕ B consists of all elements in A or
B but not both. Algorithm 11 computes the symmetric difference.

      Input : Two sets, A = {a1 , . . . , an }, B = {b1 , . . . , bm }
      Output : The symmetric difference, A ⊕ B
  1   C←∅
  2   foreach ai ∈ A do
  3      if ai 6∈ B then
  4          C ← C ∪ {ai }
  5      end
  6   end
  7   foreach bj ∈ B do
  8      if bj 6∈ A then
  9          C ← C ∪ {bj }
 10      end
 11   end
 12   output C

                 Algorithm 11: Symmetric Difference of Two Sets

Again, following the step-by-step process for analyzing this algorithm,
  1. Input: In this case, there are two sets as part of the input, A, B.
  2. Input Size: As specified, each set has cardinality n, m respectively. We could
     analyze the algorithm with respect to both input sizes, namely the input size could
     be n + m. For simplicity, to work with a single variable, we could also define
     N = n + m.
       Alternatively, we could make the following observation: without loss of generality,
       we can assume that n ≥ m (if not, switch the sets). If one input parameter is
       bounded by the other, then

                                          n + m ≤ 2n ∈ O(n)

       That is, we could simplify the analysis by only considering n as the input size.
       There will be no difference in the final asymptotic characterization as the constants
       will be ignored.
  3. Elementary Operation: In this algorithm, the most common operation is the
     set membership query (6∈). Strictly speaking, this operation may not be trivial
     depending on the type of data structure used to represent the set (it may entail a



120
                                                                            5.5 Examples

       series of O(n) comparisons for example). However, as our pseudocode is concerned,
       it is sufficient to consider it as our elementary operation.
  4. How many times is the elementary operation executed with respect to the input
     size, n? In the first for-loop (lines 2–6) the membership query is performed n times.
     In the second loop (lines 7–11), it is again performed m times. Since each of these
     loops is independent of each other, we would add these operations together to get
                                             n+m
       total membership query operations.
  5. Whether or not we consider N = n + m or n to be our input size, the algorithm is
     clearly linear with respect to the input size. Thus it is a Θ(n)-time algorithm.


5.5.3 Euclid’s GCD Algorithm

The greatest common divisor (or GCD) of two integers a, b is the largest positive integer
that divides both a and b. Finding a GCD has many useful applications and the problem
has one of the oldest known algorithmic solutions: Euclid’s Algorithm (due to the Greek
mathematician Euclid c. 300 BCE).

      Input : Two integers, a, b
      Output : The greatest common divisor, gcd(a, b)
  1   while b 6= 0 do
  2      t←b
  3      b ← a mod b
  4      a←t
  5   end
  6   Output a

                    Algorithm 12: Euclid’s GCD Algorithm

The algorithm relies on the following observation: any number that divides a, b must
also divide the remainder of a since we can write b = a · k + r where r is the remainder.
This suggests the following strategy: iteratively divide a by b and retain the remainder
r, then consider the GCD of b and r. Progress is made by observing that b and r are
necessarily smaller than a, b. Repeating this process until we have a remainder of zero
gives us the GCD because once we have that one evenly divides the other, the larger
must be the GCD. Pseudocode for Euclid’s Algorithm is provided in Algorithm 12.
The analysis of Euclid’s algorithm is seemingly straightforward. It is easy to identify the
division in line 3 as the elementary operation. But how many times is it executed with
respect to the input size? What is the input size?



                                                                                       121
5 Algorithm Analysis

When considering algorithms that primarily execute numerical operations the input is
usually a number (or in this case a pair of numbers). How big is the input of a number?
The input size of 12,142 is not 12,142. The number 12,142 has a compact representation
when we write it: it requires 5 digits to express it (in base 10). That is, the input size of
a number is the number of symbols required to represent its magnitude. Considering a
number’s input size to be equal to the number would be like considering its representation
in unary where a single symbol is used and repeated for as many times as is equal to the
number (like a prisoner marking off the days of his sentence).
Computers don’t “speak” in base-10, they speak in binary, base-2. Therefore, the input
size of a numerical input is the number of bits required to represent the number. This is
easily expressed using the base-2 logarithm function:

                                         dlog (n)e

But in the end it doesn’t really matter if we think of computers as speaking in base-10,
base-2, or any other integer base greater than or equal to 2 because as we’ve observed that
all logarithms are equivalent to within a constant factor using the change of base formula.
In fact this again demonstrates again that algorithms are an abstraction independent
of any particular platform: that the same algorithm will have the same (asymptotic)
performance whether it is performed on paper in base-10 or in a computer using binary!
Back to the analysis of Euclid’s Algorithm: how many times does the while loop get
executed? We can first observe that each iteration reduces the value of b, but by how
much? The exact number depends on the input: some inputs would only require a single
division, other inputs reduce b by a different amount on each iteration. The important
thing to realize is that we want a general characterization of this algorithm: it suffices to
consider the worst case. That is, at maximum, how many iterations are performed? The
number of iterations is maximized when the reduction in the value of b is minimized at
each iteration. We further observe that b is reduced by at least half at each iteration.
Thus, the number of iterations is maximized if we reduce b by at most half on each
iteration. So how many iterations i are required to reduce n down to 1 (ignoring the
last iteration when it is reduced to zero for the moment)? This can be expressed by the
equation:
                                           i
                                            1
                                        n       =1
                                            2

Solving for i (taking the log on either side), we get that

                                        i = log (n)

Recall that the size of the input is log (n). Thus, Euclid’s Algorithm is linear with respect
to the input size.



122
                                                                               5.5 Examples

5.5.4 Selection Sort


Recall that Selection Sort is a sorting algorithm that sorts a collection of elements by
first finding the smallest element and placing it at the beginning of the collection. It
continues by finding the smallest among the remaining n − 1 and placing it second in
the collection. It repeats until the “first” n − 1 elements are sorted, which by definition
means that the last element is where it needs to be.
The Pseudocode is presented as Algorithm 13.

      Input : A collection A = {a1 , . . . , an }
      Output : A sorted in non-decreasing order
  1   for i = 1, . . . , n − 1 do
  2      min ← ai
  3      for j = (i + 1), . . . , n do
  4         if min < aj then
  5             min ← aj
  6         end
  7         swap min, ai
  8      end
  9   end
 10   output A

                          Algorithm 13: Selection Sort
Let’s follow the prescribed outline above to analyze Selection Sort.
  1. Input: this is clearly indicated in the pseudocode of the algorithm. The input is
     the collection A.
  2. Input Size: the most natural measure of the size of a collection is its cardinality; in
     this case, n
  3. Elementary Operation: the most common operation is the comparison in line 4
     (assignments and iterations necessary for the control flow of the algorithm are not
     good candidates). Alternatively, we could have considered swaps on line 7 which
     would lead to a different characterization of the algorithm.
  4. How many times is the elementary operation executed with respect to the input
     size, n?
         • Line 4 does one comparison each time it is executed
         • Line 4 itself is executed multiple times for each iteration of the for loop in line
           3 (for j running from i + 1 up to n inclusive.



                                                                                          123
5 Algorithm Analysis

         • line 3 (and subsequent blocks of code) are executed multiple times for each
           iteration of the for loop in line 1 (for i running from 1 up to n − 1
      This gives us the following summation:
                                            n−1 X
                                            X   n
                                                            1
                                            i=1   j=i+1 |{z}

                                                  | {zline }4
                                            |       {zline 3 }
                                                   line 1

      Solving this summation gives us:
                               n−1 X
                               X   n              n−1
                                                  X
                                            1=          n−i
                                i=1 j=i+1         i=1
                                                  n−1
                                                  X          n−1
                                                             X
                                             =          n−         i
                                                  i=1        i=1
                                                                   n(n − 1)
                                             = n(n − 1) −
                                                                      2
                                                  n(n − 1)
                                             =
                                                     2


      Thus for a collection of size n, Selection Sort makes
                                                 n(n − 1)
                                                    2
      comparisons.
   5. Provide an asymptotic characterization: the function determined is clearly Θ(n2 )


5.6 Other Considerations

5.6.1 Importance of Input Size

The second step in our algorithm analysis outline is to identify the input size. Usually this
is pretty straightforward. If the input is an array or collection of elements, a reasonable
input size is the cardinality (the number of elements in the collection). This is usually
the case when one is analyzing a sorting algorithm operating on a list of elements.
Though seemingly simple, sometimes identifying the appropriate input size depends on
the nature of the input. For example, if the input is a data structure such as a graph,



124
                                                                5.6 Other Considerations

the input size could be either the number of vertices or number of edges. The most
appropriate measure then depends on the algorithm and the details of its analysis. It
may even depend on how the input is represented. Some graph algorithms have different
efficiency measures if they are represented as adjacency lists or adjacency matrices.
Yet another subtle difficulty is when the input is a single numerical value, n. In such
instances, the input size is not also n, but instead the number of symbols that would be
needed to represent n. That is, the number of digits in n or the number of bits required
to represent n. We’ll illustrate this case with a few examples.


Sieve of Eratosthenes

A common beginner mistake is made when analyzing the Sieve of Eratosthenes (named
for Eratosthenes of Cyrene, 276 BCE – 195 BCE). The Sieve is an ancient method for
prime number factorization. A brute-force algorithm, it simply tries every integer up to
a point to see if it is a factor of a given number.

      Input : An integer n
      Output : Whether n is prime or composite
  1   for i = 2, . . . , n do
  2      if i divides n then
  3          Output composite
  4      end
  5   end
  6   Output prime

                      Algorithm 14: Sieve of Eratosthenes
                                                  √                                      √
The for-loop only needs to check integers up to √    n because  any  factor greater than  n
would necessarily have a corresponding
                             √           factor
                                          √       n. A naive approach would observe that
the for-loop gets executed n − 1 ∈ O( n) times which would lead to the (incorrect)
impression that the Sieve is a polynomial-time (in fact sub-linear!) running time algorithm.
Amazing that we had a primality testing algorithm over 2,000 years ago! In fact, primality
testing was a problem that was not known to have a deterministic polynomial time
running algorithm until 2001 (the AKS Algorithm [1]).
The careful observer would realize that though n is the input, the actual input size is
again log (n), the number of bits required to represent n. Let N = log (n) be a placeholder
for our actual input size (and so n = 2N ). Then the running time of the Sieve is actually
                                       √          √
                                     O( n) = O( 2N )

which is exponential with respect to the input size N .



                                                                                       125
5 Algorithm Analysis


  1   int a = 45, m = 67;
  2   int result = 1;
  3   for(int i=1; i<=n; i++) {
  4     result = (result * a % m);
  5   }

                        Code Sample 5.8: Naive Exponentiation


This distinction is subtle but crucial: the difference between a polynomial-time algorithm
and an exponential algorithm is huge even for modestly sized inputs. What may take
a few milliseconds using a polynomial time algorithm may take billions and billions of
years with an exponential time algorithm as we’ll see with our next example.


Computing an Exponent

As a final example, consider the problem of computing a modular exponent. That is,
given integers a, n, and m, we want to compute

                                           an mod m

A naive (but common!) solution might be similar to the Java code snippet in Code
Sample 5.8.
Whether one chooses to treat multiplication or integer division as the elementary operation,
the for-loop executes exactly n times. For “small” values of n this may not present a
problem. However, for even moderately large values of n, say n ≈ 2256 , the performance
of this code will be terrible.
To illustrate, suppose that we run this code on a 14.561 petaFLOP (14 quadrillion floating
point operations per second) super computer cluster (this throughput was achieved by
the Folding@Home distributed computing project in 2013). Even with this power, to
make 2256 floating point operations would take

                                 2256
                              15
                                                       ≈ 2.5199 × 1053
                   14.561 × 10 · 60 · 60 · 24 · 365.25
or 252 sexdecilliion years to compute!
For context, this sort of operation is performed by millions of computers around the
world every second of the day. A 256-bit number is not really all that “large”. The
problem again lies in the failure to recognize the difference between an input, n, and
the input’s size. The real solution to this problem is an algorithm known as Repeated
Squaring where values are squared,

                          (a2 )2 = a4 , (a4 )2 = a8 , (a8 )2 = a1 6, . . .



126
                                                                 5.6 Other Considerations


  1    public static double average(double arr[]) {
  2      double sum = 0.0;
  3      for(int i=0; i<arr.length; i++) {
  4        sum = sum + arr[i];
  5      }
  6      return sum / arr.length;
  7    }

                          Code Sample 5.9: Computing an Average

allowing the exponent to grow exponentially and reducing the number of total multipli-
cations.


5.6.2 Control Structures are Not Elementary Operations

When considering which operation to select as the elementary operation, we usually
do not count operations that are necessary to the control structure of the algorithm.
For example, assignment operations, or operations that are necessary to execute a loop
(incrementing an index variable, a comparison to check the termination condition).
To see why, consider the method in Code Sample 5.9. This method computes a simple
average of an array of double variables. Consider some of the minute operations that
are performed in this method:
      • An assignment of a value to a variable (lines 2, 3, and 4)
      • An increment of the index variable i (line 3)
      • A comparison (line 3) to determine if the loop should terminate or continue
      • The addition (line 4) and division (line 6)
A proper analysis would use the addition in line 4 as the elementary operation, leading
to a Θ(n) algorithm. However, for the sake of argument, let’s perform a detailed analysis
with respect to each of these operations. Assume that there are n values in the array,
thus the for loop executes n times. This gives us
      • n + 2 total assignment operations
      • n increment operations
      • n comparisons
      • n additions and
      • 1 division
Now, suppose that each of these operations take time t1 , t2 , t3 , t4 , and t5 milliseconds



                                                                                        127
5 Algorithm Analysis

each (or whatever time scale you like). In total, we have a running time of
      t1 (n + 2) + t2 n + t3 n + t4 n + t5 = (t1 + t2 + t3 + t4 )n + (2t1 + t5 ) = cn + d
Which doesn’t change the asymptotic complexity of the algorithm: considering the
additional operations necessary for the control structures only changed the constant
sitting out front (as well as some additive terms).
The amount of resources (in this case time) that are expended for the assignment,
increment and comparison for the loop control structure are proportional to the true
elementary operation (addition). Thus, it is sufficient to simply consider the most
common or most expensive operation in an algorithm. The extra resources for the control
structures end up only contributing constants which are ultimately ignored when an
asymptotic analysis is performed.


5.6.3 Average Case Analysis

The behavior of some algorithms may depend on the nature of the input rather than
simply the input size. From this perspective we can analyze an algorithm with respect
to its best-, average-, and worst-case running time.
In general, we prefer to consider the worst-case running time when comparing algorithms.


Example: searching an array

Consider the problem of searching an array for a particular element (the array is unsorted,
contains n elements). You could get lucky (best-case) and find it immediately in the first
index, requiring only a single comparison. On the other hand, you could be unlucky and
find the element in the last position or not at all. In either case n comparisons would be
required (worst-case).
What about the average case? A naive approach would be to average the worst and best
case to get an average of n+1
                            2
                               comparisons. A more rigorous approach would be to define
a probability of a successful search p and the probability of an unsuccessful search, 1 − p.
For a successful search, we further define a uniform probability distribution on finding
the element in each of the n indices. That is, we will find the element at index i with
probability np . Finding the element at index i requires i comparisons. Thus the total
number of expected comparisons in a successful search is
                                    n
                                    X          p   p(n + 1)
                                          i·     =
                                    i=1
                                               n      2

For an unsuccessful search, we would require n comparisons, thus the number of expected
comparisons would be
                                        n(1 − p)



128
                                                              5.6 Other Considerations

                                     p      C
                                     0      n

                                     1    7    1
                                            n+
                                     4    8    8
                                     1    3    1
                                            n+
                                     2    4    4
                                     3    5    3
                                            n+
                                     4    8    8
                                          n+1
                                     1
                                           2
Table 5.6: Expected number of comparisons C for various values of the probability of a
           successful search p.

Since these are mutually exclusive events, we sum the probabilities:
                          p(n + 1)              p − pn + 2n
                                   + n(1 − p) =
                             2                       2
We cannot remove the p terms, but we can make some observations for various values
(see Table 5.6.3). When p = 1 for example, we have the same conclusion as the naive
approach. As p decreases, the expected number of comparisons grows to n.




   Figure 5.3: Expected number of comparisons for various success probabilities p.



5.6.4 Amortized Analysis

Sometimes it is useful to analyze an algorithm not based on its worst-case running time,
but on its expected running time. On average, how many resources does an algorithm



                                                                                     129
5 Algorithm Analysis

use? This is a more practical approach to analysis. Worst-case analysis assumes that all
inputs will be difficult, but in practice, difficult inputs may be rare. The average input
may require fewer resources.
Amortized algorithm analysis is similar to the average-case analysis we performed before,
but focuses on how the cost of operations may change over the course of an algorithm.
This is similar to a loan from a bank. Suppose the loan is taken out for $1,000 at a 5%
interest rate. You don’t actually end up paying $50 the first year. You pay slightly less
than that since you are (presumably) making monthly payments, reducing the balance
and thus reducing the amount of interest accrued each month.
We will not go into great detail here, but as an example, consider Heap Sort. This
algorithm uses a Heap data structure. It essentially works by inserting elements into the
heap and then removing them one by one. Due to the nature of a heap data structure,
the elements will come out in the desired order.
An amortized analysis of Heap Sort may work as follows. First, a heap requires about
log (n) comparisons to insert an element (as well as remove an element, though we’ll
focus on insertion), where n is the size of the heap (the number of elements currently in
the heap. As the heap grows (and eventually shrinks), the cost of inserts will change.
With an initially empty heap, there will only be 0 comparisons. When the heap has
1 element, there will be about log (1) comparisons, then log (2) comparisons and so on
until we insert the last element. This gives us
                                                n−1
                                                X
                                     C(n) =           log (i)
                                                i=1

total comparisons.


5.7 Analysis of Recursive Algorithms

Recursive algorithms can be analyzed with the same basic 5 step process. However,
because they are recursive, the analysis (step 4) may involve setting up a recurrence
relation in order to characterize how many times the elementary operation is executed.
Though cliche, as a simple example consider a naive recursive algorithm to compute the
n-th Fibonacci number. Recall that the Fibonacci sequence is defined as

                                     Fn = Fn−1 + Fn−2

with initial conditions F0 = F1 = 1. That is, the n-th Fibonacci number is defined as the
sum of the two previous numbers in the sequence. This defines the sequence

                              1, 1, 2, 3, 5, 8, 13, 21, 34, 55, . . .

The typical recursive solution is presented as Algorithm 15



130
                                                               5.7 Analysis of Recursive Algorithms


         Input : An integer n
         Output : The n-th Fibonacci Number, Fn
     1   if n = 0 or n = 1 then
     2       output 1
     3   else
     4       output Fibonacci(n − 1) + Fibonacci(n − 2)
     5   end

                                 Algorithm 15: Fibonacci(n)


The elementary operation is clearly the addition in line 4. However, how do we analyze
the number of additions performed by a call to Fibonacci(n)? To do this, we setup
a recurrence relation. Let A(n) be a function that counts the number of additions
performed by Fibonacci(n). If n is zero or one, the number of additions is zero (the base
case of our recursion performs no additions). That is, A(0) = A(1) = 0. But what about
n > 1?
When n > 1, line 4 executes. Line 4 contains one addition. However, it also contains
two recursive calls. How many additions are performed by a call to Fibonacci(n − 1)
and Fibonacci(n − 2)? We defined A(n) to be the number of additions on a call to
Fibonacci(n), so we can reuse this function: the number of additions is A(n − 1) and
A(n − 2) respectively. Thus, the total number of additions is

                                    A(n) = A(n − 1) + A(n − 2) + 1

This is a recurrence relation,7 in particular, it is a second-order linear non-homogeneous
recurrence relation. This particular relation can be solved. That is, A(n) can be expressed
as a non-recursive closed-form solution.
The techniques required for solving these type of recurrence relations are beyond the
scope of this text. However, for many common recursive algorithms, we can use a simple
tool to characterize their running time, the “Master Theorem.”


5.7.1 The Master Theorem

Suppose that we have a recursive algorithm that takes an input of size n. The recursion
may work by dividing the problem into a subproblems each of size nb (where a, b are
constants). The algorithm may also perform some amount of work before or after the
recursion. Suppose that we can characterize this amount of work by a polynomial
function, f (n) ∈ Θ(nd ).
 7
     Also called a difference equation, which are sort of discrete analogs of differential equations.




                                                                                                        131
5 Algorithm Analysis

This kind of recursive algorithm is common among “divide-and-conquer” style algorithms
that divide a problem into subproblems and conquers each subproblem. Depending on
the values of a, b, d we can categorize the runtime of this algorithm using the Master
Theorem.
Theorem 4 (Master Theorem). Let T (n) be a monotonically increasing function that
satisfies
                          T (n) = aT ( nb ) + f (n)
                          T (1) = c
where a ≥ 1, b ≥ 2, c > 0. If f (n) ∈ Θ(nd ) where d ≥ 0, then
                                     
                                      Θ(nd )        if a < bd
                           T (n) =     Θ(nd log n) if a = bd
                                       Θ(nlogb a )   if a > bd
                                     


“Master” is a bit of a misnomer. The Master Theorem can only be applied to recurrence
 relations of the particular form described. It can’t, for example, be applied to the previous
 recurrence relation that we derived for the Fibonacci algorithm. However, it can be used
 for several common recursive algorithms.


Example: Binary Search

As an example, consider the Binary Search algorithm, presented as Algorithm 16. Recall
that binary search takes a sorted array (random access is required) and searches for an
element by checking the middle element m. If the element being searched for is larger
than m, a recursive search on the upper half of the list is performed, otherwise a recursive
search on the lower half is performed.
Let’s analyze this algorithm with respect to the number of comparisons it performs. To
simplify, though we technically make two comparisons in the pseudocode (lines 5 and 7),
let’s count it as a single comparison. In practice a comparator pattern would be used and
logic would branch based on the result. However, there would still only be one invocation
of the comparator function/object. Let C(n) be a function that equals the number of
comparisons made by Algorithm 16 while searching an array of size n in the worst case
(that is, we never find the element or it ends up being the last element we check).
As mentioned, we make one comparison (line 5, 7) and then make one recursive call. The
recursion roughly cuts the size of the array in half, resulting in an array of size n2 . Thus,
                                              n
                                    C(n) = C        +1
                                                2

Applying the master theorem, we find that a = 1, b = 2. In this case, f (n) = 1 which is
bounded by a polynomial: a polynomial of degree d = 0. Since
                                      1 = a = bd = 20



132
                                                              5.7 Analysis of Recursive Algorithms


      Input    : A sorted collection of elements A = {a1 , . . . , an }, bounds 1 ≤ l, h ≤ n,
                 and a key ek
      Output : An element a ∈ A such that a = ek according to some criteria; φ if no
                 such element exists
  1   if l > h then
  2       output φ
  3   end
  4   m ← b h+l
              2
                c
  5   if am = ek then
  6       output am
  7   else if am < ek then
  8       BinarySearch(A, m + 1, h, e)
  9   else
 10       BinarySearch(A, l, m − 1, e)
 11   end

                      Algorithm 16: Binary Search – Recursive


by case 2 of the Master Theorem applies and we have

                                          C(n) ∈ Θ(log (n))


Example: Merge Sort

Recall that Merge Sort is an algorithm that works by recursively splitting an array of
size n into two equal parts of size roughly n2 . The recursion continues until the array
is trivially sorted (size 0 or 1). Following the recursion back up, each subarray half is
sorted and they need to be merged. This basic divide and conquer approach is presented
in Algorithm 17. We omit the details of the merge operation on line 4, but observe that
it can be achieved with roughly n − 1 comparisons in the worst case.

      Input : An array, sub-indices 1 ≤ l, r ≤ n
      Output : An array A0 such that A[l, . . . , r] is sorted
  1   if l < r then
  2       MergeSort(A, l, b r+l
                             2
                                c)
  3      MergeSort(A, d r+l
                         2
                            e, r)
  4      Merge sorted lists A[l, . . . , b r+l
                                            2
                                               c] andA[d r+l
                                                          2
                                                             e, . . . , r]
  5   end

                               Algorithm 17: Merge Sort



                                                                                              133
5 Algorithm Analysis

Let C(n) be the number of comparisons made by Merge Sort. The main algorithm makes
two recursive calls on subarrays of size n2 . There are also n + 1 comparisons made after
the recursion. Thus we have
                                               n
                                  C(n) = 2C          + (n − 1)
                                                 2
Again, applying the Master Theorem we have that a = 2, b = 2 and that f (n) = (n + 1) ∈
Θ(n) and so d = 1. Thus,
                                  2 = a = bd = 21
and by case 2,
                                     C(n) ∈ Θ(n log (n))

We will apply the Master Theorem to several other recursive algorithms later on.


5.8 Exercises

When asked to design and analyze an algorithm, be sure to provide the following:
   1. Complete pseudocode
   2. Identify the input and the input size, n
   3. Identify the elementary operation
   4. Compute how many times the elementary operation is executed with respect to
      the input size n
   5. Provide a Big-O asymptotic characterization for the algorithm’s complexity
Exercise 5.1. Let P be an image represented as an (n × m) 2-dimensional array of
pixels. Design an algorithm that given an image P will rotate it clockwise by 90 degrees.
Exercise 5.2. Let C be a set of circles each represented as a triple (x, y, r) where x, y is
its center and r is its radius. Design and analyze an algorithm that given a set C of n
circles determines if any of the circles intersect.
Exercise 5.3. Let A = [a1 , a2 , . . . , an ] be a collection of integers. A pair (i, j) is called
an inversion if i < j but ai > aj . For example, if A = [2, 3, 8, 6, 1] then the list of
inversions is (1, 5), (2, 5), (3, 4), (3, 5), (4, 5). Design an algorithm that, given a collection
of integers A outputs a list of its inversions.

For the next few questions, consider defining a time function with respect to n. That is,
if the algorithm is linear, we could write the time that it takes as a function of n:

                                         t = f (n) = cn



134
                                                                           5.8 Exercises

If it is quadratic, we could write it as a quadratic function:

                                     t = f (n) = cn2

Both of these instances ignore lower order terms. If we know the time it takes for a
particular value of n then we can compute the constant c and consequently predict the
time t it takes for other values of n or vice versa.
Exercise 5.4. An algorithm takes 1.5 ms for an input size 100. How long will it take
for input size 1500 (assuming that low-order terms are negligible) if the running time is
  1. linear
  2. O(n log n)
  3. quadratic
  4. cubic
  5. exponential
Exercise 5.5. An algorithm takes 0.5 ms for input size 100. How large can an input size
be if a problem can be solved in 1 minute (assuming that low-order terms are negligible)
if the running time is:
  1. linear
  2. O(n log n)
  3. quadratic
  4. cubic
  5. exponential
Exercise 5.6. Prove each of the following statements by applying the definition of Big-O.
That is, derive an inequality (show your work) and clearly identify the c, n0 constants
you derive as per the definition of Big-O.

  1. 3n = O(n)
        √
  2. 500 n = O(n)
  3. n2 + 2n + 1 = O(n2 )
  4. 2048n + 1234 = O(n2 )
  5. n log (32n) = O(n log (n)))
  6. 12n3 + 50n2 − 12n − 60 = O(n3 )
  7. n2n = O(3n )
  8. log (n!) = O(n log n)



                                                                                     135
6 Trees

6.1 Introduction

One of our fundamental goals is to design a data structure to store elements that offers
efficient and arbitrary retrieval (search), insertion, and deletion operations. We’ve already
seen several examples of list-based data structures that offer different properties. Array-
based lists offer fast index-based retrieval and, if sorted even more efficient, O(log n)
key-based retrieval using binary search. However, arbitrary insertion and deletion
may result in shifts leading to O(n) behavior. In contrast, linked lists offer efficient,
O(1) insertion and deletion at the head/tail of the list, but inefficient, O(n) arbitrary
search/insert/delete.
Stacks and queues are efficient in that their core functionality offers O(1) operations
(push/pop, enqueue/dequeue) but they are restricted-access data structures and do not
offer arbitrary and efficient operations.
We how turn our attention to another fundamental data structure based on trees. Trees
are a very special type of graph. Graphs are useful because they can model many types
of relations and processes in physics, biology, finance, and pretty much every other
discipline. Moreover, we have over two centuries of foundational work on graph results
and algorithms to work with.1 General graphs, however, are not necessarily as structured
as they need to be for an efficient data structure. Trees, however, are highly structured
and have several useful properties that we can exploit to design an efficient data structure.
As we will see, trees have the potential to offer efficient O(log n) behavior for all three
fundamental operations.


6.2 Definitions & Terminology

A graph is a collection of nodes such that pairs of nodes may be connected by edges.
More formally,
Definition 7 (Undirected Graph). A graph is a two-tuple, G = (V, E) where
      • V = {v1 , . . . , vn } is a set of vertices
 1
     Graph Theory usually credited to Euler who studied the Seven Bridges of Königsberg [6] in the 18th
      century.




                                                                                                    137
6 Trees


                                                               b
                                     a




                                                 d                 e
                               c



                                                           g
                                           f

                     Figure 6.1: An undirected graph with labeled vertices.

      • E ⊆ V × V is a set of edges connecting nodes

In general, the edges connecting two vertices may be directed or undirected. However,
we’ll only focus on undirected edges so that if two vertices u, v are connected, then the
edge that connects them can be written as an unordered pair, e = (u, v) = (v, u). An
example of a graph can be found in Figure 6.1.
In the example, the nodes have been labeled with single characters. In general, we’ll
want to store elements of any type in a graph’s vertices. In the context of data structures
when vertices hold elements, they are usually referred to as nodes and we’ll use this
terminology from here on. Another point to highlight is that the size of a graph is usually
measured in terms of the number of nodes in it. In the definition we have designated the
variable n to denote this. In general, a graph may have up to
                                     
                                      n      n(n − 1)
                                          =
                                      2         2
edges connecting pairs of nodes.2


Trees

General graphs may model very complex binary relations. Other than the graph structure
itself (nodes possibly connected by edges) there is not much structure to exploit. For
this reason, we will instead focus on a subclass of graphs called trees.
     This notation is from combinatorics, nk is a binomial coefficient and can be read as “n choose k.”
 2
                                               

      This operation counts the number of ways there are   to choose an unordered subset of k elements
      from a set of n unique elements. In this case, n2 is counting the number of ways there are two
      choose a pair of vertices from a set of n vertices.




138
                                                           6.2 Definitions & Terminology

Definition 8 (Trees). A tree is an acyclic graph.

A cycle in a graph is any sequence of pairwise connected nodes that begin and end at
the same node. For example, the sequence of vertices aegcda in the graph in Figure 6.1
forms a cycle. Likewise, the sequences begdb, bgdb, and cdgc also form cycles as well as
many others. A graph in which there are no cycles is referred to as an acyclic graph and
has a much simpler structure. This simplicity gives a tree several properties that can be
exploited. An example of several trees can be found in Figure 6.2.
The example in Figure 6.2(d) is actually a disconnected tree in that there are several
components whose nodes are not connected by any edges. Such graphs can be seen as a
collection of trees and are usually called forests. Though forests are useful in modeling
some problems, we will again restrict our attention to connected trees such that there is
only one connected component.


Structural Properties

By definition, trees are acyclic which already gives them a high degree of structure.
There are several related properties that follow from this structure. First, since we are
restricting attention to connected trees, it will always be the case that the number of
edges in a tree will be equal to one less than the number of vertices. That is:

Lemma 7. In any tree,
                                       |E| = n − 1
where n = |V |.

Recall that in general, graphs may have as many as O(n2 ) vertices. However, trees will
necessarily have fewer O(n) edges. This makes trees sparse graphs. Many graph problems
that are hard or do not admit themselves to efficient solutions become easy or even trivial
on sparse graphs and trees in particular.
Another consequence of a lack of cycles means that any two nodes in a tree will be
connected by exactly one, unique path. A path is much like a cycle except that it need
not begin/end at the same node. In Figure 6.2(a), the sequence of vertices af cgb form a
path. This is, in fact, the only path connecting the vertices a and b. Furthermore, the
length of a path is defined as the number of edges on it. The aforementioned path af cgb
has length 4. Note that the length of.a path is also equal to the number of nodes in the
path minus one. In the same tree in Figure 6.2(a), af cgcf ad is also technically a path,
however, it traverses several edges more than once, backtracking from g. Such a path is
referred to as a non-simple path. A path that does not traverse an edge more than once
is a simple path and we will restrict our consideration to simple paths.

Lemma 8. Let T = (V, E) be a connected tree and let u, v ∈ V be two nodes in T .
Then there exists exactly one single unique path connecting u, v.



                                                                                       139
6 Trees




                                                                           a



                                           b                   b           c               d
                  a


                                                               e           f               g

                               d               e
          c
                                                               h            i              j


                                       g
                       f                                                   k
      (a) The previous graph from Figure 6.1 with           (b) Another tree drawn in a more
      enough edges removed to make it a tree.               organized manner.

                                                                   c
                                                       b
                                                                           d
              a            b          c

                                                                                       h
                                                              a
              d            e

                                                                           e

              f                                    f                               g
          (c) Yet another little tree, hap-    (d) A disconnected tree with two components.
          pier.

Figure 6.2: Several examples of trees. It doesn’t matter how the tree is depicted or
            organized, only that it is acyclic. The final example, 6.2(d) represents a
            disconnected tree, called a forest.




140
                                                              6.2 Definitions & Terminology


                                               b


                      a                        g                            g


           d          e        f               c                       b         c


                               c               f                                 f


                               g               a                                 a


                               b          d         e                       d         e
            (a) Tree rooted at a.     (b) Tree rooted at b.        (c) Tree rooted at g.

       Figure 6.3: Several possible rooted orientations for the tree from Figure 6.2(a).


Proof. First we observe that there has to be at least one path between u, v. Otherwise,
if there were not, then T would not be a connected tree.
Now suppose, by way of contradiction that there exist at least two paths p and p0
connecting u, v such that p = 6 p0 . However, this means that we can now form a cycle.
Starting at u and taking the first path p to v and then from v returning to u via p0 . Since
this forms a cycle, it contradicts the fact that we started with a tree. Thus, there cannot
be two distinct paths between two vertices.


These properties will prove useful in adapting trees as collection data structures.


Orienting Trees

To give trees even more structure, we can orient them. The first way that we’ll do this
is by rooting them. Imagine that you could take a tree (say any tree in Figure 6.2) and
lift it from the page3 by grabbing a single node. Then shake the tree and let gravity
dangle the remaining nodes downward. Then place the tree back onto the page. The
node that you shook the tree from will be designated as the tree’s root and all other
nodes are oriented downward. An example can be found in Figure 6.3 where we have
done this operation on the tree in Figure 6.2(a) by dangling from various nodes.
 3
     Or screen if you will.




                                                                                           141
6 Trees

The alert reader and professional arborists4 alike will notice that we’ve drawn our trees
with the root at the top and its “branches” growing downward, the exact opposite of how
most real, organic trees grow. Trees in Computer Science “grow” downward because that
is how we read. Since trees will ultimately be used to store data it is easier for a human
to read them top-to-bottom, left-to-right. Of course, when represented in a computer,
there is no such orientation; it is only a convention that we humans use when drawing
and describing them. This convention is quite old; in fact Arthur Cayley, the original
mathematician who first studied trees drew them like this in his original paper [3].5
The vertical orientation of a trees lends itself to some obvious terminology, borrowed
from the concept of a family tree. Let u be a tree node. The node immediately above it
is called its parent. Any node(s) immediately below it are called u’s children. Likewise,
all nodes on the path from u all the way back up to the root are known as u’s ancestors;
u’s children and all nodes connected below them are called u’s descendants. Suppose we
were to remove all nodes in a tree except u and u’s descendant(s). The new tree would
form a subtree rooted at u. Other tree-related terminology will be used. For example,
the root of the tree will be the only node without a parent. If a node has no children, it
will be referred to as a leaf node.


Binary Trees

We can place more useful restrictions on our tree. An oriented tree such that every node
has at most two children is a binary tree. This means that every node in a binary tree
may have 0 children (a leaf), 1 child, or 2 children. We can further horizontally orient
the children in a binary tree. Since the number of children is restricted to at most 2, we
can refer to them as a left child and right child. All the identified relations in a binary
tree node are depicted in Figure 6.4. Since a node may only have 1 child, using this
orientation means that it may have a left child and missing its right child, or it may have
a right child and missing its left child.
A larger binary tree is depicted in Figure 6.5. In this tree, the nodes k, l, n, o, p, q, r are
all leaf nodes. Many nodes have both a left and right child (b, g, j as examples) while
some have a left child, but no right child (m, i for example) and some have a right child,
but no left child (e, h for example).
Given this orientation, it is natural to define the depth of a node. Let u be a node in
a binary tree with root r. The depth of u is the length of the unique path from r to u.
The depth of the root, r itself is 0 by convention. The depth of a tree T itself is defined
as the maximal depth (i.e. “deepest”) of any node in the tree. Thus, if a tree consists
of a single root node, its depth is zero.6 . It is possible to have an empty tree (where
|V | = 0) and so by convention and for consistency, the depth of an empty tree is usually
 4
   Or dendrologists.
 5
   Specifically in part LVIII. On the Analytical Forms called Trees.–Part II in which he essentially defines
    binary trees.
 6
   The author proposes to call this a seed ; let’s hope it catches on.




142
                                                                 6.2 Definitions & Terminology




                                                            ancestors

                             Parent
                                                        p


                                                u
                Left Child                                         Right Child

                                        `               r



                                       descendants



Figure 6.4: The various relations of a tree node. For a given node u, its parent, left
            and right child are labelled. In addition, nodes above u can collectively be
            referred to as ancestors (including its parent) and nodes in the subtree at
            u can collectively be referred to as descendants (including its immediate
            children).



                                                    a


                                 b                                    c


                        d               e                                         f

                    g        h                  i                         j           k

                l       m         n         o                     p           q


                    r

                                 Figure 6.5: A Binary Tree




                                                                                          143
6 Trees


                         Table 6.1: Depths of nodes in Figure 6.5
                                      Node(s)       Depth
                                          a           0
                                         b, c         1
                                       d, e, f        2
                                     g, h, i, j, k    3
                                   l, m, n, o, p, q   4
                                          r           5




      Figure 6.6: A complete tree of depth d = 3 which has 1 + 2 + 4 + 8 = 15 nodes.




−1. The depth of every node in the tree in Figure 6.5 is given in Table 6.1. Since r is
the deepest node, the depth of the tree itself is 5.
When depicting trees, we draw all nodes of equal depth on the same horizontal line. All
nodes on this line are said to be at the same level in the tree. Levels are numbered as
the depth so the root is at level 0, the next depth nodes are at level 1, etc. Structurally,
we could draw the tree however we want. However, placing nodes of a different depth at
different levels makes the tree less readable.
Now, consider a complete (also called full or perfect) binary tree of depth d. That is, a
binary tree in which all nodes are present at all levels; from level 0 up to level d, the
depth of the tree. How many nodes total are there in such a tree? At level 0, there would
only be the root. At level 1, there would be 2 children from the root. Since all nodes are
present, at level 2, both children from the previous 2 nodes would be present giving a
total of 4, etc. That is, at each level, the number of nodes on that level doubles from the
previous level. A concrete example for d = 3 is given in Figure 6.6 which has a total of
15 nodes.
We can generalize this by adding up powers of 2. At level 0, there are 20 nodes, level
1, there are 21 nodes, level 2 there are 22 nodes, etc. Summing all of these nodes up to



144
                                                                     6.2 Definitions & Terminology




                                                            Level        Number of Nodes
                                                              0                    20

                                                              1                    21

                                                              2                    22

                                                              3                    23

                                                              4                    24

                                                              ..                   ..
                                                               .                    .

                                                              d                    2d
                                                                          d
                                                                          X
                                                            total               i = 2d+1 − 1
                                                                          i=0




Figure 6.7: A summation of nodes at each level of a complete binary tree up to depth d.


depth d gives us the following geometric series which has a well-known solution.7
                                                                   d
                                                                   X
                     n = 20 + 21 + 22 + · · · + 2d−1 + 2d =              2k = 2d+1 − 1
                                                                   k=0

This is visualized in Figure 6.7.
Taken from another perspective, if we have a complete binary tree with n nodes, its
depth is
                                d = log (n + 1) − 1
That is, the depth is logarithmic, d ∈ O(log n). This observation gives us our motivation
for using binary trees as a collection data structure. If we can develop a tree-based data
structure such that insertion, retrieval, and removal operations are all proportional to
 7
     As a Computer Scientist, you should also find it familiar–it represents the maximum integer repre-
      sentable by d bits.




                                                                                                  145
6 Trees

the depth d then we have the potential for a very efficient, generalized collection data
structure. Of course it is not necessarily a simple matter as we will see shortly.


6.3 Implementation

A binary tree implementation is fairly straightforward and similar to that of a linked
list. Each tree node holds references to its two children, its parent (which is optional,
but simplifies a lot of operations) and the key element stored in the tree node. In
pseudocode we’ll use the same dot operator syntax as with a linked list, so that each of
these components can be referenced as follows. Let u be a tree node, then
      • u.key is its key,
      • u.parent is its parent,
      • u.lef tChild is its left child, and
      • u.rightChild is its right child.
The tree itself only needs a reference to the root element, represented using T.root. Note
that we do not keep a reference to every single tree node. Keeping track of and updating
every tree node would bring us right back to a linked list or array-based list to store all
the nodes, defeating the purpose of exploiting the tree structure. We could, however,
keep track of how many nodes are in the tree as this could easily be updated with each
insert/remove operation just as with a linked list.
As a concrete example, a pair of Java classes may be written to implement these two
objects. In particular, the tree node class may look something like the following (getters,
setters and other standard methods are omitted).

  1    public class BinaryTreeNode<T> {
  2

  3        private   T key;
  4        private   TreeNode<T> parent;
  5        private   TreeNode<T> leftChild;
  6        private   TreeNode<T> rightChild;
  7

  8        ...
  9    }


While a tree itself would look something like the following.

  1    public class BinaryTree<T> {
  2

  3        private TreeNode<T> root;



146
                                                                         6.4 Tree Traversal


  4        private int size;
  5

  6        ...
  7    }



6.4 Tree Traversal

Before we develop the basic operations that will give us a collection data structure we
need to ensure that we have a way to process all the elements in a general binary tree.
With a linked list, we could easily iterate over the elements stored in it by starting at
the head and iterating over each element until the tail. With a binary tree the order in
which we iterate over elements is not all that clear. It is natural to start at the root
node, but where do we go from there? To the left child or the right? If we go to the left,
then we must remember to eventually come back and iterate over the right child.
More generally given a node u and its children, u.lef tChild, u.rightChild in which order
do we enumerate them? In fact, there are several tree traversal strategies each with
its own advantages and applications. Each one traverses the three different nodes in a
different order but all are based on a more general graph traversal algorithm called DFS
which attempts to explore a graph as deeply as possible before backtracking to explore
the other areas. In general we will prefer to descend left in the tree before we backtrack
and explore the right subtree. There is nothing particularly special about this preference
(other than reading left-to-right is more natural). We could prefer to to right before
going left. However, this would be equivalent to going left while also “reversing” the tree
(reversing the left/right child at each node).
To understand the traversal strategies better, we will present the traversal choices locally.
At any particular node, u, we’ll refer to this node as the root (think of the subtree rooted
at u) and its left and right child respectively. Given that we’ll always go left-before-right,
this gives three possible orders in which to enumerate the three nodes:
      • Preorder: root-left-right
      • Inorder: left-root-right
      • Postorder: left-right-root
Of course, any given node may be missing one or both of its children in which case we
do not enumerate it, but otherwise, we will preserve these orderings.
When we describe a traversal strategy, it is not restricted to merely enumerating all the
elements in a tree. The purpose of using a binary tree as a data structure is to store
data in each of its nodes. When visiting a node, in general we can perform any operation
or process on the data stored in it rather than simply enumerating the element. The
details of how to process a node would be specific to the type of data being stored and



                                                                                         147
6 Trees

                                                a


                                     b                      c


                                 d        e                     f

                              Figure 6.8: A small binary tree.


the overall goal of the algorithm being executed.




6.4.1 Preorder Traversal

A preorder traversal enumerates nodes in a root-left-right manner. This means that
we process a node immediately when we first visit it, then after processing it, we next
process its left child. Only after we have processed its left child and all of its descendants,
do we return to process the right child and its descendants. However, we need a way
to “remember” that the right child and its subtree still need to be traversed. As a small
example, consider the tree in Figure 6.8. Starting at the root, we enumerate a. We next
visit its left subtree rooted at b. When we visit b, we immediately enumerate it and again
traverse left and enumerate its left child, d. Since d has no children, we backtrack to b.
Having already enumerated b and its left child d, we next traverse to its right child, e.
We enumerate e and since it has no children, we backtrack all the way back up to the
root a and continue to its right child, c. We immediately enumerate c, but since it has
no left child to visit, we traverse to its right child f as the final node. Altogether, the
order of enumeration was

                                         a, b, d, e, c, f


We can implement a preorder traversal algorithm using a stack to “remember” each right
child that we need to visit and what order to visit them. Initially, we push the tree’s root
onto the stack. We then go into a loop that executes until the stack is emptied and we
have processed all the nodes in the tree. On each iteration, we pop the node on the top
of the stack and process it. We then need to process its children before we continue with
other elements on the stack. To set up the next iteration, we push these child nodes onto
the stack so that they are the next ones to be processed. We need to take care to exploit
the LIFO ordering of the stack properly, however. We want the left child to be processed
before the right child, so we push the right child first, then the left child second. The full



148
                                                                        6.4 Tree Traversal

      push a           push m           (enter loop)        print i            push q
                        push l         pop, node = h                           push p
   (enter loop)        print g            push n         (enter loop)          print j
  pop, node = a                          (no push)      pop, node = o
      push c          (enter loop)        print h         (no push)          (enter loop)
      push b         pop, node = l                        (no push)         pop, node = p
     print a           (no push)       (enter loop)         print o           (no push)
                       (no push)      pop, node = n                           (no push)
   (enter loop)          print l        (no push)        (enter loop)          print p
  pop, node = b                         (no push)       pop, node = c
      push e         (enter loop)        print n           push f            (enter loop)
      push d        pop, node = m                         (no push)         pop, node = q
      print b         (no push)         (enter loop)        print c           (no push)
                        push r         pop, node = e                          (no push)
   (enter loop)        print m             push i        (enter loop)           print q
  pop, node = d                          (no push)      pop, node = f
     push h           (enter loop)         print e         push k            (enter loop)
      push g         pop, node = r                          push j          pop, node = k
     print d           (no push)        (enter loop)       print f            (no push)
                       (no push)       pop, node = i                          (no push)
   (enter loop)          print r         (no push)       (enter loop)          print k
  pop, node = g                            push o       pop, node = j


      Figure 6.9: A walkthrough of a preorder traversal on the tree from Figure 6.5.


traversal is presented as Algorithm 18.

      Input : A binary tree, T
      Output : A preorder traversal of the nodes in T
  1   S ← empty stack
  2   push T.root onto S
  3   while S is not empty do
  4      u ← S.pop
  5      process u
  6      push u.rightChild onto S
  7      push u.lef tChild onto S
  8   end

   Algorithm 18: Stack-based Preorder Tree Traversal. If a node does not exist,
   we implicitly do not push it onto the stack.
A full preorder traversal on the binary tree in Figure 6.5 would result in the following
ordering of vertices.
                         a, b, d, g, l, m, r, h, n, e, i, o, c, f, j, p, q, k
A full walkthrough of the algorithm on this example is presented in Figure 6.9.



                                                                                         149
6 Trees

A common alternative way of presenting a preorder traversal is to use recursion. This
alternative version is presented as Algorithm 19. We are still implicitly using a stack
to keep track of where we’ve come from in the tree, but instead of using a stack data
structure, we are exploiting the system call stack to do this.

      Input : A binary tree node u
      Output : A preorder traversal of the nodes in the subtree rooted at u
  1   if u = null then
  2       return
  3   else
  4       process u
  5       preOrderTraversal(u.lef tChild)
  6       preOrderTraversal(u.rightChild)
  7   end

   Algorithm 19: preOrderTraversal(u): Recursive Preorder Tree Traversal



Applications

A preorder traversal is straightforward and simple to implement. It is the most common
traversal strategy for many applications that do not require tree nodes to be traversed
in any particular order. It can be used to build a tree, enumerate elements in a tree
collection, copy a tree, etc.



6.4.2 Inorder Traversal

An inorder traversal strategy visits nodes in a left-root-right order. This means that
instead of immediately processing a node the first time we visit it, we wait until we have
processed its left child and all of its descendants before processing it and moving on to
the right subtree. From another we process a node when we backtrack to it (after having
processed the left subtree.
Again, consider the small tree in Figure 6.8. We again start at the root a, but do not
immediately enumerate it. Instead, we traverse to its left child, b. However, again we
do not enumerate b until after we have enumerated its left subtree. We traverse to d
and since there is no left child, we enumerate d and return to b. Now that we have
enumerated b’s left subtree, we now enumerate b and traverse down to its right child e
and enumerate it. We now backtrack all the way back up to a. Since we have enumerated
its entire left subtree, we can now enumerate a and continue to its right subtree. Since c
has no left child, c is enumerated next, and then f . Altogether, the order of enumeration



150
                                                                                 6.4 Tree Traversal

was

                                           d, b, e, a, c, f


We can use the same basic idea of using a stack to keep track of tree nodes, however we
want to delay processing the node until we’ve explored the left subtree. To do this, we
need a way to tell if we are visiting the node for the first time or returning from exploring
the left subtree (so that we may process it). To achieve this, we allow the node to be null
as a flag to indicate that we are backtracking from the left subtree. Thus, if the current
node is not null, we push it back onto the stack for later processing and explore the left
subtree. If it is null, we pop the node on the top of the stack and process it, then push
its right child to setup the next iteration. The full process is described in Algorithm 21.

      Input : A binary tree, T
      Output : An inorder traversal of the nodes in T
  1   S ← empty stack
  2   u ← T.root
  3   while S is not empty Or u 6= null do
  4      if u 6= null then
  5          push u onto S
  6          u ← u.lef tChild
  7      else
  8         u ← S.pop
  9         process u
 10         u ← u.rightChild
 11      end
 12   end

               Algorithm 20: Stack-based Inorder Tree Traversal
In lines 6 and 10, if no such child exists, u becomes null, indicating that we are backtracking
on the next iteration.
A full inorder traversal on the binary tree in Figure 6.5 would result in the following
order of vertices.

                          l, g, r, m, d, h, n, b, e, o, i, a, c, p, j, q, f, k

Again, a full walkthrough of the algorithm on this example is presented in Figure 6.10.
As with a preorder traversal, we can use recursion to implicitly use the call stack to keep
track of the ordering. The only difference is is when we process the given node. In the
preorder case, we did it before the two recursive calls. In the inorder case, we process it



                                                                                               151
6 Trees




(enter loop, u = a)     (enter loop,                             u = null)
     push a              u = null)         (enter loop,        pop i, update           (enter loop,
   update u = b        pop r, update        u = null)              u=i                  u = null)
                           u=r            pop b, update          process i            pop j, update
(enter loop, u = b)      process r            u=b             update u = null             u=j
     push b           update u = null       process b                                   process j
   update u = d                           update u = e            (enter loop,        update u = q
                         (enter loop,                              u = null)
(enter loop, u = d)       u = null)     (enter loop, u = e)     pop a, update       (enter loop, u = q)
     push d            pop m, update          push e                 u=a                  push q
   update u = g             u=m          update u = null           process a         update u = null
                          process m                              update u = c
(enter loop, u = g)   update u = null       (enter loop,                              (enter loop,
     push g                                  u = null)        (enter loop, u = c)      u = null)
   update u = l         (enter loop,      pop e, update             push c           pop q, update
                         u = null)             u=e             update u = null           u=q
(enter loop, u = l)    pop d, update         process e                                 process q
     push l                u=d             update u = i          (enter loop,       update u = null
 update u = null         process d                                u = null)
                       update u = h     (enter loop, u = i)     pop c, update          (enter loop,
   (enter loop,                               push i                u=c                 u = null)
    u = null)       (enter loop, u = h)    update u = o           process c           pop f , update
  pop l, update           push h                                update u = f              u=f
       u=l           update u = null (enter loop, u = o)                                process f
     process l                              push o       (enter loop, u = f )         update u = k
 update u = null        (enter loop,    update u = null        push f
                         u = null)                          update u = j      (enter loop, u = k)
    (enter loop,      pop h, update       (enter loop,                              push k
     u = null)             u=h             u = null)     (enter loop, u = j) update u = null
  pop g, update          process h       pop o, update         push j
        u=g            update u = n          u=o            update u = p          (enter loop,
     process g                             process o                               u = null)
   update u = m     (enter loop, u = n) update u = null (enter loop, u = p)     pop k, update
                          push n                               push p                u=k
(enter loop, u = m) update u = null       (enter loop,    update u = null          process k
      push m                               u = null)                           update u = null
   update u = r         (enter loop,     pop i, update       (enter loop,
                         u = null)           u=i              u = null)              (done)
(enter loop, u = r)   pop n, update        process i       pop p, update
      push r               u=n          update u = null         u=p
 update u = null         process n                            process p
                     update u = null      (enter loop,    update u = null


      Figure 6.10: A walkthrough of a inorder traversal on the tree from Figure 6.5.




152
                                                                       6.4 Tree Traversal

between them. The recursive version is presented as Algorithm 21.

      Input : A binary tree node u
      Output : An inorder traversal of the nodes in the subtree rooted at u
  1   if u = null then
  2       return
  3   else
  4       inOrderTraversal(u.lef tChild)
  5       process u
  6       inOrderTraversal(u.rightChild)
  7   end

      Algorithm 21: inOrderTraversal(u): Recursive Inorder Tree Traversal


Applications

The primary application of an inorder tree traversal is from where it derives its name. If
we perform an inorder traversal on a binary search tree (see the next section), then we
get an enumeration of elements that is sorted or “in order.”


6.4.3 Postorder Traversal

The final depth-first-search based traversal strategy is a postorder traversal in which
nodes are visited in a left-right-root manner. That is, we hold off on processing a node
until both of its children and all of their respective descendants have been processed.
We turn once again to the small binary tree in Figure 6.8. We start at the root a and
descend all the way to the left and process d since it has no children. Backtracking to
b we again hold off on processing it until we’ve traversed its right child, e. Only after
we’ve enumerated both d and e do we finally enumerate b. At this point, a’s left subtree
has been enumerated, but again we hold off on processing a until we are done with its
right subtree. c has no left child, so we enumerate its right child, f first and only then
do we process c. Finally, the last node to be processed is the root itself, a.
We can again utilize a stack data structure to perform a postorder traversal, but it gets
a bit more complicated as we need to distinguish 3 different cases. When visiting a node,
we need to know if it is the first time we’ve traversed it, the second time (returning from
its left subtree) or the third time (returning from its right subtree, thus it needs to be
processed). To do this, we keep track of not only the current node but also the previous
node to know where we came from, either the parent, the left child, or the right child.
That way, we can tell if it is the first time we’ve visited the node (the previous node
would be the parent), the second time (the previous node is the left child) or the third



                                                                                      153
6 Trees

time (the previous node is the right child). The full postorder traversal is presented in
Algorithm 22.

      Input : A binary tree, T
      Output : A postorder traversal of the nodes in T
  1   S ← empty stack
  2   prev ← null
  3   push T.root onto S
  4   while S is not empty do
  5      curr ← S.peek
  6      if prev = null Or prev.lef tChild = curr Or prev.rightChild = curr then
  7          if curr.lef tChild 6= null then
  8              push curr.lef tChild onto S
  9          else if curr.rightChild 6= null then
 10              push curr.rightChild onto S
 11          end
 12       else if curr.lef tChild = prev then
 13          if curr.rightChild 6= null then
 14              push curr.rightChild onto S
 15          end
 16       else
 17          process curr
 18          S.pop
 19       end
 20       prev ← curr
 21   end

              Algorithm 22: Stack-based Postorder Tree Traversal
A full postorder traversal on the binary tree in Figure 6.5 would result in the following
order of vertices.
                         l, r, m, g, n, h, d, o, i, e, b, p, q, j, k, f, c, a
Again, a full walkthrough of the algorithm on this example is presented in Figure 6.11.

      prev = null
        push a                (enter loop)              (enter loop)              (enter loop)
                           update curr = (b)         update curr = (d)         update curr = (g)
      (enter loop)                check:                    check:                    check:
   update curr = (a)    prev.lef tChild = curr    prev.lef tChild = curr    prev.lef tChild = curr
  check: prev = null     ((b).lef tChild = (b))    ((d).lef tChild = (d))    ((g).lef tChild = (g))
       push (b)                 push (d)                  push (g)                  push (l)
    update prev = a         update prev = b           update prev = d           update prev = g




154
                                                                                 6.4 Tree Traversal

                                  (enter loop)                                       (enter loop)
      (enter loop)            update curr = (m)            (enter loop)           update curr = (o)
   update curr = (l)                  check:            update curr = (n)                check:
          check:           prev.rightChild = curr              check:          prev.lef tChild = curr
prev.lef tChild = curr     (null.rightChild = (m)) prev.rightChild = curr       ((o).lef tChild = (o))
 ((l).lef tChild = (l))               check:        (null.rightChild = (n))             (noop)
         (noop)             curr.lef tChild = prev           process n             update prev = o
    update prev = l          ((r).lef tChild = (r))      update prev = n
                               update prev = m                                       (enter loop)
      (enter loop)                                         (enter loop)           update curr = (o)
   update curr = (l)              (enter loop)          update curr = (h)               check:
         check:               update curr = (m)                check:          prev.rightChild = curr
prev.rightChild = curr                check:        prev.rightChild = curr     (null.rightChild = (o))
(null.rightChild = (l))    prev.rightChild = curr (null.rightChild = (h))              process o
        process l          (null.rightChild = (m))           process h             update prev = o
    update prev = l                process m             update prev = h
                               update prev = m                                        (enter loop)
       (enter loop)                                        (enter loop)            update curr = (i)
   update curr = (g)              (enter loop)          update curr = (d)                 check:
           check:              update curr = (g)               check:          prev.rightChild = curr
prev.rightChild = curr                check:        prev.rightChild = curr     (null.rightChild = (i))
(null.rightChild = (g))    prev.rightChild = curr ((n).rightChild = (d))                  check:
           check:          (null.rightChild = (g))           process d          curr.lef tChild = prev
 curr.lef tChild = prev             process g            update prev = d         ((o).lef tChild = (o))
  ((l).lef tChild = (l))        update prev = g                                     update prev = i
        push (m)                                           (enter loop)
    update prev = g               (enter loop)          update curr = (b)            (enter loop)
                               update curr = (d)               check:             update curr = (i)
      (enter loop)                    check:        prev.rightChild = curr              check:
  update curr = (m)        prev.rightChild = curr ((h).rightChild = (b))       prev.rightChild = curr
         check:             ((m).rightChild = (d))             check:          (null.rightChild = (i))
prev.rightChild = curr                check:         curr.lef tChild = prev            process i
((m).rightChild = (m))      curr.lef tChild = prev    ((d).lef tChild = (d))       update prev = i
       push (r)              ((g).lef tChild = (g))          push (e)
   update prev = m                  push (h)             update prev = b             (enter loop)
                                update prev = d                                   update curr = (e)
      (enter loop)                                         (enter loop)                 check:
   update curr = (r)              (enter loop)          update curr = (e)      prev.rightChild = curr
          check:               update curr = (h)               check:          (null.rightChild = (e))
prev.lef tChild = curr                check:        prev.rightChild = curr             process e
 ((r).lef tChild = (r))    prev.rightChild = curr ((e).rightChild = (e))           update prev = e
         (noop)             ((h).rightChild = (h))           push (i)
    update prev = r                 push (n)             update prev = e              (enter loop)
                                update prev = h                                   update curr = (b)
      (enter loop)                                         (enter loop)                  check:
   update curr = (r)              (enter loop)          update curr = (i)      prev.rightChild = curr
         check:                update curr = (n)               check:           ((i).rightChild = (b))
prev.rightChild = curr                check:        prev.rightChild = curr             process b
(null.rightChild = (r))    prev.rightChild = curr    ((i).rightChild = (i))        update prev = b
        process r           ((n).rightChild = (n))           push (o)
    update prev = r                  (noop)              update prev = i           (enter loop)
                                update prev = n                                  update curr = (a)




                                                                                                   155
6 Trees

           check:             update curr = (p)             (enter loop)
prev.rightChild = curr               check:              update curr = (q)             (enter loop)
 ((e).rightChild = (a))    prev.lef tChild = curr               check:              update curr = (k)
           check:           ((p).lef tChild = (p))    prev.rightChild = curr              check:
 curr.lef tChild = prev             (noop)            (null.rightChild = (q))    prev.rightChild = curr
  ((b).lef tChild = (b))       update prev = p                process q          (null.rightChild = (k))
         push (c)                                         update prev = q                process k
    update prev = a              (enter loop)                                        update prev = k
                              update curr = (p)             (enter loop)
       (enter loop)                  check:              update curr = (j)             (enter loop)
   update curr = (c)       prev.rightChild = curr              check:               update curr = (f )
          check:           (null.rightChild = (p))    prev.rightChild = curr              check:
prev.rightChild = curr             process p          (null.rightChild = (j))    prev.rightChild = curr
 ((c).rightChild = (c))        update prev = p                process j          (null.rightChild = (f ))
        push (f )                                         update prev = j                process f
    update prev = c               (enter loop)                                       update prev = f
                               update curr = (j)             (enter loop)
       (enter loop)                   check:              update curr = (f )            (enter loop)
   update curr = (f )      prev.rightChild = curr                check:             update curr = (c)
          check:           (null.rightChild = (j))    prev.rightChild = curr               check:
prev.rightChild = curr                check:           ((q).rightChild = (f ))   prev.rightChild = curr
((f ).rightChild = (f ))    curr.lef tChild = prev               check:           ((k).rightChild = (c))
        push (j)             ((p).lef tChild = (p))    curr.lef tChild = prev             process c
    update prev = f                 push (q)            ((j).lef tChild = (j))       update prev = c
                                update prev = j                push (k)
      (enter loop)                                         update prev = f              (enter loop)
   update curr = (j)              (enter loop)                                      update curr = (a)
          check:              update curr = (q)             (enter loop)                   check:
prev.lef tChild = curr                check:             update curr = (k)       prev.rightChild = curr
 ((j).lef tChild = (j))    prev.rightChild = curr               check:           ((f ).rightChild = (a))
        push (p)            ((q).rightChild = (q))    prev.rightChild = curr             process a
    update prev = j                  (noop)           ((k).rightChild = (k))         update prev = a
                               update prev = q                 (noop)
       (enter loop)                                       update prev = k




      Figure 6.11: A walkthrough of a postorder traversal on the tree from Figure 6.5.




A recursive version would follow the same pattern as before. With a postorder traversal,
we would process the node after both of the recursive calls. A recursive postorder traversal



156
                                                                        6.4 Tree Traversal

is presented in Algorithm 23.

       Input : A binary tree node u
       Output : A postorder traversal of the nodes in the subtree rooted at u
   1   if u = null then
   2       return
   3   else
   4       postOrderTraversal(u.lef tChild)
   5       postOrderTraversal(u.rightChild)
   6       process u
   7   end

  Algorithm 23: postOrderTraversal(u): Recursive Postorder Tree Traversal


Applications

    • Topological sorting
    • Destroying a tree when manual memory management is necessary (roots are the
      last thing that get cleaned up)
    • Reverse polish notation (operand-operand-operator, unambiguous, used in old HP
      calculators)
    • PostScript (Page Description Language)


6.4.4 Tree Walk Traversal

 It turns out that we can perform any of the three depth-first-search based tree traversals
without using any extra space at all (that is, a stack is not necessary) by exploiting the
 oriented structure of a binary tree. As with the postorder traversal, we simply need
 to keep track of where we came from (a previous node) and where we are (a current
 node) to determine where to go next. By only keeping track of two variables, we can
 also determine when a node should be processed. This traversal is generally called a
“tree walk” and it is illustrated on the small tree example we’ve been using in Figure 6.12.
The traversal is essentially a “walk” around the perimeter of the tree.
The rules that we follow are quite simple and only require local information. In particular,
there are only three cases that we need to distinguish. Suppose we are at a node u as
our current node with parent p, and left/right child as `, r respectively as in Figure 6.13.
The rules are outlined in Table 6.2.
The full tree walk is presented as Algorithm 24 and includes all three traversal strategies.



                                                                                       157
6 Trees




                                              a


                                    b                  c


                               d        e                  f

                  Figure 6.12: A tree walk on the tree from Figure 6.8.




                                   preorder
                                                  p


                                        u         postorder



                               `                  r

                                   inorder

      Figure 6.13: The three general cases of when to process a node in a tree walk.




158
                                                                         6.4 Tree Traversal


                              Table 6.2: Rules for Tree Walk
            If previous is... Then traverse to... And process u if order is...
                   p                   `                   Preorder
                   `                   r                     Inorder
                   r                   p                   Postorder


                                                a


                                       b                 c


                                  d        e                 f

                      Figure 6.14: A Breadth First Search Example



The algorithm actually only keeps track of the previous node’s type (whether it was a
parent, left, or right child) rather than the node itself. It is a lot more complex than the
three simple rules in Table 6.2 because we need to take care of many corner cases where
the left and/or right children may not exist.




6.4.5 Breadth-First Search Traversal

An alternative to the depth-first-search based traversal strategies, we can explore a binary
tree using what is known as a Breadth First Search (BFS) traversal. BFS is also a general
graph traversal algorithm that explores a graph by visiting the closest vertices first before
venturing deeper into the graph. When applied to an oriented binary tree, BFS ends up
exploring a graph in a top-to-bottom, left-to-right ordering. In this manner, all nodes at
the same depth are enumerated before moving on to the next deepest level. The basic
enumeration is depicted in Figure 6.14 which represents a BFS traversal on the same
graph we’ve been using in previous examples. Unsurprisingly, the order of enumeration
matches the labeling which was chosen to be top-bottom/left-right for readability.
The visualization of the ordering should give a hint as to the implementation of BFS. If
we were to stretch out the ordering it would be clear that the nodes are enumerated in
one straight long line, suggesting a queue should be used. Indeed, we start by enqueueing
the root node. Then on each iteration, we dequeue the next node and enqueue its children
to be processed later. Since a queue is FIFO, we enqueue in the left child first. This is



                                                                                         159
6 Trees


      Input : A binary tree, T
      Output : A Tree Walk around T
  1   curr ← T.root
  2   prevT ype ← parent
  3   while curr 6=null do
  4      if prevT ype = parent then
             //preorder: process curr here
  5          if curr.lef tChild exists then
                 //Go to the left child:
  6              curr ← curr.lef tChild
  7              prevT ype ← parent
  8          else
  9             curr ← curr
 10             prevT ype ← lef t
 11       else if prevT ype = lef t then
             //inorder: process curr here
 12          if curr.rightChild exists then
                 //Go to the right child:
 13              curr ← curr.rightChild
 14              prevT ype ← parent
 15          else
 16             curr ← curr
 17             prevT ype ← right
 18       else if prevT ype = right then
             //postorder: process curr here
 19          if curr.parent = null then
                 //root has no parent, we’re done traversing
 20              curr ← curr.parent
            //are we at the parent’s left or right child?
 21         else if curr = curr.parent.lef tChild then
 22            curr ← curr.parent
 23            prevT ype ← lef t
 24          else
 25             curr ← curr.parent
 26             prevT ype ← right
 27   end

                 Algorithm 24: Tree Walk based Tree Traversal



160
                                                                   6.5 Binary Search Trees

presented as Algorithm 25.

       Input : A binary tree, T
       Output : A BFS traversal of the nodes in T
   1   Q ← empty queue
   2   enqueue T.root into Q
   3   while Q is not empty do
   4      u ← Q.dequeue
   5      process u
   6      enqueue u.lef tChild into Q
   7      enqueue u.rightChild into Q
   8   end

                 Algorithm 25: Queue-based BFS Tree Traversal
Contrast the BFS algorithm with the original stack-based preorder algorithm (Algorithm
18). In fact, they are nearly identical. The only difference is that we use a queue instead
of a stack (and its corresponding operations) and reverse the operations on the left/right
child. This calls back to one of the major themes of this text on the importance of
“smart” data structures. Simply by changing the underlying data structure used in an
algorithm, it results in a fundamentally different but complementarily useful property.


6.5 Binary Search Trees

Binary trees have a lot of structure but there is still not enough to make them efficient
data structures. As we’ve presented them so far, we can store elements and retrieve them,
but without any additional ordered structure, using any one of the general traversal
strategies we’ve developed so far is still going to be O(n) in the worst case. We will now
add some addition structure to our binary trees in order to attempt to make the general
operations of insertion, retrieval and deletion more efficient. In particular, it will be our
goal to make all three of these operations have complexity that is proportional to the
depth of the tree, O(d). To do that, we introduce Binary Search Trees (BSTs).
Definition 9 (Binary Search Tree). A binary search tree is a binary tree such that every
node u has an associated key, u.key which satisfies the binary search tree property:
   1. Every node in the left subtree of u has a key value less than u.key
   2. Every node in the right subtree of u has a key value greater than u.key

Implicitly, this definition does not allow for duplicate key values in a binary tree. In
general, we could amend this definition to allow duplicate keys but we would need to be
consistent on if we place duplicates in the left or right subtree. In any case, it is not that



                                                                                          161
6 Trees

big of a deal. In general, we can “break ties” using some unique value to each objects to
induce a total ordering on all elements stored in a binary search tree.8
Furthermore, though we will use integer key values in all of our examples, in practice
binary search trees need not be restricted to such values as they are intended to hold
any type of object or data that we wish to store. Typically, an implementation will use
a hash value of an object or data as a key value. A hash value is simply the result of
applying a hash function that maps an object to an integer value. This is such a common
operation it is built into many programming languages.

                                                                   50

                                                  40                     60

                                        20              45                          90

                                  10         24               49              80         95

                              2        15          28    48             75         85

                                  12


                                   Figure 6.15: A Binary Search Tree

The binary search tree property is an inductive property. Since it holds for all nodes, it
follows that every rooted subtree in a BST is also a binary search tree. A full example
can be found in Figure 6.15. Note that a binary search tree need not be full or complete
as in the example. Many children (and thus entire subtrees) are not present. The binary
search tree property merely guarantees that all keys in the less subtree are less and those
in the right subtree are greater than the key stored in the root respectively.


6.5.1 Retrieval

We can exploit the binary search tree property similar to how binary search exploits a
sorted array. With binary search, we check the middle element and either find what we
were searching for or we have effectively cut the array in half as we know that we should
either search in the lower half or the upper half. Similarly, if with a binary search tree,
we start a search for a particular key k at the root. If the root’s key value is greater
than k (k < T.root.key) then we know that the key lies in its left subtree while if the
search key value is greater (T.root.key < k) then we know that it must lie in the right
subtree. In either case, we continue the search on the tree rooted at the left or right
child respectively. If we ever find a node such that the key value matches we can stop
 8
     In practice, all things being equal, we could use the memory address of two objects stored in a tree to
      break ties.




162
                                                                  6.5 Binary Search Trees

the search and return the node’s value. Of course, not every search will necessarily be
successful. If we search for a key value that does not exist in the tree, we will eventually
reach the end of the tree at some leaf node. If this happens, we can stop the process
and return a flag indicating an unsuccessful search. Several examples are illustrated in
Figure 6.16.
In some of the examples we got lucky and did not have to search too deeply in the
tree. In fact, if we had searched for the key value k = 50 only a single key comparison
would have been required. Thus, in the best case only O(1) comparisons are necessary
for a search/retrieval operation. In the worst case, however, we had to search to the
deepest part of the tree as in the case where we successfully searched for k = 12 and
unsuccessfully searched for k = 13. In both cases we made 6 key comparisons before
ending the search. In general, we may need to make d comparisons where d is the depth
of the tree. This achieves what we set out to do as a retrieval operation is O(d). The full
search process is presented as Algorithm 26.

      Input : A binary search tree, T , a key k
      Output : The tree node u ∈ T such that u.key = k, φ if no such node exists.
  1   u ← T.root
  2   while u 6= φ do
  3      if u.key = k then
  4          output u
  5      else if u.key > k then
  6          u ← u.lef tChild
  7      else if u.key < k then
  8          u ← u.rightChild
  9      end
 10   end
 11   output φ

            Algorithm 26: Search algorithm for a binary search tree



6.5.2 Insertion

Inserting a new node into a binary search tree is a simple extension to searching for a
node. In fact, it is exactly the same process. Since we will not allow duplicate key values,
if we perform a search and find that a node with the given key value already exists, we
can make the same design decisions we used with lists (throw an exception, return an
error flag, or treat it as a no op). Otherwise, we perform the same search and when
we reach the end of the tree, we insert the node at the spot where we would otherwise
expect it to be.



                                                                                        163
6 Trees




                                        50                                                                 50

                       40                     60                                          40                     60

             20              45                          90                     20              45                          90

       10         24               49              80         95          10         24               49              80         95

   2        15          28    48             75         85            2        15          28    48             75         85

       12                                                                 12

(a) Searching a BST Example 1. In this example,                    (b) Searching a BST Example 2. In this example,
we search for the key value k = 45. The search                     we search for the key value k = 12. The search
starts at the root. Since 45 < 50 we traverse to the               starts at the root. Since 12 < 50 we traverse to
left child. The next comparison finds that 40 < 45                 the left child. The search then traverses left twice
so we traverse right. Finally, the last comparison                 more. When compared to 10, it traverses right and
finds that the key value matches at which point                    finally left, finding the key value at the deepest
the search process terminates.                                     leaf in the tree. A total of 6 comparisons was
                                                                   necessary.

                                        50                                                                 50

                       40                     60                                          40                     60

             20              45                          90                     20              45                          90

       10         24               49              80         95          10         24               49              80         95

   2        15          28    48             75         85            2        15          28    48             75         85

       12                                                                 12




(c) Searching a BST Example 3. In this example,                    (d) Searching a BST Example 4. Another unsuc-
we search for the key value k = 55 which will result               cessful search when we search for the key value
in an unsuccessful search. Starting at the root,                   k = 13. Similar to example 6.16(b). With the
we traverse right, then left at which point we are                 same 6 comparisons, we’ve traversed off the tree
no longer in the tree. With only 2 comparisons,                    and return our flag value.
we are able to conclude and return a flag value
indicating an unsuccessful search.

                   Figure 6.16: Various Search Examples on a Binary Search Tree




164
                                                                        6.5 Binary Search Trees

                                                  15

                                          7                   30

                                  4           9         20         40

                              1       5            17        23

                                                  16

Figure 6.17: Insertion of a new node into a binary search tree. A search is performed for
             the key k = 16, when the open slot is found, the node is inserted.


This is illustrated in Figures 6.16(c) and 6.16(d) where a dashed “phantom” node is
indicated where the key values would otherwise have been. When performing a search we
will always be guaranteed to be inserting the new node as a leaf node. We’ll never insert
a new node in the middle of the tree so no other considerations are necessary. Once
we have performed the search, creating a new node and inserting it only requires a few
reference shuffles. Thus, using the same analysis as before, in the worst case, insertion is
also O(d). Developing the pseudocode for the insertion operation is left as an exercise
(see Exercise 6.16). A small example is demonstrated in Figure 6.17.


6.5.3 Removal

The removal of keys also follows a similar initial search operation. Given a key k to
delete, we traverse the binary search tree for the node containing k which is an O(d)
operation. Suppose the node we wish to delete is u. There are several cases to take care
of depending on how many children u has.
Case 1: Suppose that u is a leaf node and has no children. It is a simple matter to
delete it by simply changing its parent node’s reference. Though we do need to check if
u is the left or right child. In particular,
   • If u = u.parent.lef tChild then set u.parent.lef tChild ← φ; otherwise
   • If u = u.parent.rightChild then set u.parent.rightChild ← φ.
Case 2: Suppose that u has only one child (left or right) and is missing the other child.
We don’t want to “prune” the subtree of its child, but at the same time we want to
preserve the binary search tree property. To do this we simply “promote” the only child
up to u’s position. Specifically, suppose that u’s child is a left child. The steps to to
promote it are nearly exactly the same as used to delete (circumvent) a node in a linked
list.
If u is a left child then:



                                                                                           165
6 Trees

   • u.parent.lef tChild ← u.lef tChild
   • u.lef tChild.parent ← u.parent
Otherwise if u is a right child then:
   • u.parent.rightChild ← u.rightChild
   • u.rightChild.parent ← u.parent
The operations for the case that u only has a right child is analogous. An example of
this operation is depicted in Figures 6.18(c) and 6.18(d).
Case 3: Suppose that u has both of its children. Our first instinct may be to promote
one of its children up to its place similar to the previous case. In some instances this
might work. For example, consider the tree in Figure 6.15. If we wanted to delete the
root, 50, we could promote its right child, 60 up to its place. However, that is only
because 60 has no left child. If it had a left child (say 55), then it would not be possible
as promoting 60 up would mean it would then have to have 3 children (40, 55, 90).
There are many potential solutions to this problem, but we want to ensure that the
operation, in addition to preserving the binary search tree property, is simple, efficient,
and causes a minimal change to the tree’s structure. Instead of immediately deleting
the node, let us instead remove the key, resulting in an “empty” node. We will want to
backfill this empty node with a key value in the tree that preserves the BST property.
Again, we exploit the binary search tree property. Because the subtree rooted at u is also
a BST, every key in its left subtree is less than its key and every key in its right subtree
is greater. Thus, there are two candidates that we could “promote”: the maximum value
in the left subtree or the minimum value in the right subtree. Suppose we go with the
first option, By replacing u with the the maximum value in the left subtree, we still
ensure that all elements in the left subtree are less than it (because it was the maximum
value) and all nodes in the right subtree are still greater than it (since it was less than u
to begin with).
There are two issues that we need to deal with, however. One is finding the maximum
(or minimum) value in a subtree and the other is how do we delete the node containing
the maximum value from the tree so that it can take the place of u? We’ll deal with the
second problem first. Once we’ve found the node containing the maximum, if it has zero
or two children, we already know how to deal with that from the two previous cases.
The only case we really have to deal with is if it has two children. However, we only have
to think for a moment that this is, in fact, impossible! Suppose that a node, x contained
the maximum key value and it had both children. By the binary search tree property,
its right child, x.rightChild necessarily has a key value greater than i, thus x cannot
contain the maximum key value! That is, necessarily the maximum must only have at
most 1 child and we know how to deal with that situation.
How do we actually find the maximum value? Again, we exploit the binary search tree
property. We first traverse to the left child. From there, we only have to keep traversing



166
                                                                    6.5 Binary Search Trees

right until there is no longer a right child. Conversely, to find the minimum value in the
right subtree, we traverse to the right child and then keep traversing left until there is no
left child. The process for the former choice is presented as Algorithm 27.

       Input : A node u in a binary search tree with two children.
       Output : The node containing the maximum key value in u’s left subtree.
  1    curr ← u.lef tChildt
  2    while curr.rightChild 6= φ do
  3       curr ← curr.rightChild
  4    end
  5    output curr

      Algorithm 27: Finding the maximum key value in a node’s left subtree.
Examples of cases 2 and 3 can be found in Figure 6.18. The full delete operation has
involves several subroutines. Finding the node to be deleted (or determining it does not
exist) is O(d). Finding the minimum in the left subtree is also at most O(d). Ultimately,
swapping the keys and/or references is O(1). Since these are all independent operations,
we have the total complexity as O(d) + O(d) + O(1) = O(d).



6.5.4 In Practice

We achieved our goal of developing a tree-based data structure that offered O(d) insert,
retrieval, and update operations. However, we have still fallen short of our main goal
of ensuring that these operations are all O(log n). The reason for this is that a binary
search tree may become skewed or degenerate. Consider the binary search tree that would
be formed if we inserted the elements 10, 20, 30, . . . , 100 in that order. The resulting tree
would look that that in Figure 6.19. This tree is degenerate because its depth is linear
with respect to the number of nodes in the tree; d = n − 1 ∈ O(n). This should look
familiar: a degenerate binary search tree is essentially a linked list!
In practice trees will not always be degenerate. However, there is no guarantee that
the depth with be logarithmic. Yet more structure would be necessary to make this
guarantee. There are plenty of examples of balanced binary search trees such that insert
and remove operations reorder or rebalance a tree so that for every node the left/right
subtrees are roughly equal in depth, ensuring that d ∈ O(log n). Each one has its own
advantages and complexity in operations such as AVL trees, 2-3/B-trees, Red-Black trees,
splay trees, and treaps. We will not examine such data structures here, but know that it
is possible to guarantee that the depth, and thus the basic operations are all efficient,
O(log n).
Many programming languages also offer standard implementations for binary search
trees, in particular balanced BSTs. Java, for example, provides a TreeSet (as well as a



                                                                                          167
6 Trees



                                             15                                                                        9

                                7                               30                                      7                             30

                    4                    9             20             40                       4                             20             40

            1               5                     17        23                        1            5                    17        23
          (a) Deletion of a node with two children                              (b) Node 15 is replaced with the ex-
          (15). First step: find the maximum node                               tremal node, preserving the BST prop-
          in the left-sub-tree (lesser elements).                               erty

                                                  9                                                               9

                                    7                            30                                4                              30

                        4                                  20             40              1                  5              20             40

                1               5                  17           23                                                     17        23
           (c) Deletion a node with only one child                                   (d) Removal is achieved by simply
           (7).                                                                      promoting the single child/subtree.

Figure 6.18: BST Deletion Operation Examples. Figures 6.18(a) and 6.18(b) depict the
             deletion of a node (15) with two children. Figures 6.18(c) and 6.18(d) depict
             the deletion of a node with only one child (7).



                                             10

                                                      20

                                                            30

                                                                     40

                                                                           50

                                                                                60

                                                                                          70

                                                                                                   80

                                                                                                            90

                                                                                                                 100



                                        Figure 6.19: A degenerate binary search tree.




168
                                                                                                  6.6 Heaps

 TreeMap ) which is a red-black tree implementation with guaranteed O(log n) operations.
Moreover, the default iterator pattern is an inorder traversal, providing a sorted ordering.
It is usually used with one of its interfaces, a SortedSet which allows you to use a
 Comparator for ordering.


6.6 Heaps

We may have fallen short of presenting a guaranteed efficient general tree data structure,
however we will present a related data structure that does provide efficiency guarantees.
Like stack and queues, however, this data structure’s efficiency will come at the cost of
restricting access to its elements.

Definition 10 (Heap). A heap is a binary tree of depth d that satisfies the following
properties.
  1. It is a full up to level d − 1. That is, every node is present at every level up to and
     including level d − 1.
  2. At level d all nodes are full-to-the-left. That is, all nodes at the deepest level are
     all as far left as possible.
  3. It satisfies the heap property; every node has a key value that is greater than both
     of its children.

This definition actually defines what is referred to as a max-heap. This is because as a
consequence of the heap property, the maximum element is always guaranteed to be at
the root of the heap. Alternatively, we could redefine a heap so that every node’s children
has key values that are greater than it. This would define a min-heap. In practice the
distinction is unimportant and a comparator is usually used to define order. Thus, you
can get a min-heap implementation using a max-heap with a comparator that reverses
the usual ordering. A larger min-heap example can be found in Figure 6.20.

                                                           1

                                        5                                     30

                              50                 55                 40                  80

                        60         53       72        70       65        45        90        95

                   92        63




                                    Figure 6.20: A min-heap



                                                                                                       169
6 Trees

As depicted it is easy to see the fullness properties. Every node is present at every level
except the last one. In the deepest level, the two remaining nodes are as far left as
possible (the children of 60). Essentially there are no “gaps” in any of the levels. This
fullness property guarantees that a heap’s depth will always be logarithmic, O(log n).
However, it is easy to see that a heap does not provide too much additional structure on
the ordering of elements other than the fact that the maximum element is at the top of
the heap. In fact, the structure of a heap resembles a real heap: it is an unorganized pile
of stuff. If we were to throw something on top of a real heap, we may not have much
control on where it ends up. The operations on a heap data structure are similar. We
cannot perform efficient arbitrary searches on a heap because of this lack of structure.
We can, however, support two core restricted access operations. We can add elements to
the heap and we can remove the top most (getMax) element.


6.6.1 Operations

To develop the two core operations, we’ll assume that we are working with a max-heap as
in Definition 10. We’ll use the smaller max-heap example in Figure 6.21 in our examples.

                                                 65

                                       32                  50

                                 25         18        27        8

                            10        23


                                 Figure 6.21: A Max-heap



Insert Key

The first operation is to insert a new key into the heap. Our first instinct may be to start
at the root as we did with binary search trees. We could then move down left/right in
the heap in some manner to insert a new key, but how? Suppose we were to insert a new
key, k = 90 into the max-heap in Figure 6.21. Clearly it would need to become the new
root of the heap, displacing 65. We could continue to shove 65 down displacing either its
left or right child. However, what criteria would we use? With only local information to
go on, we may choose to exchange with the larger of the two children. Following this
criteria we would then exchange 65 and 50, 50 and 27, resulting in something that looks
like Figure 6.22.
The problem is that this is not a valid heap as it does not fulfill the fullness property.



170
                                                                              6.6 Heaps

                                                90

                                      32                   65

                                25         18         50        8

                           10        23          27


                           Figure 6.22: An Invalid Max-heap


One might be tempted to modify this approach and instead always exchange the inserted
node with the lesser of the two children. This strategy would work with this particular
example, but we could easily come up with a counter example in which it fails (in fact,
the same example, just swap the keys 32 and 50 and it will still fail).
Let’s take a step back and redevelop an approach that first ensures that the fullness
property is satisfied but the heap property need not necessarily be preserved (at first
at least). Preserving the fullness property means that we would necessarily insert the
key at the deepest level at the left-most available spot. For the example in Figure 6.21,
this would mean inserting as the left child of 18. Inserting 90 in this manner clearly
violates the heap heap property, however, so we need to “fix” the heap. This process is
known as heapifying. We exchange the inserted key with its parent until either 1) the
heap property is satisfied or 2) we’ve reached the root node and the inserted key has
become the new root of the heap. This process is fully illustrated in Figure 6.23.
The heapify process is presented as Algorithm 28 which assumes that the new key has
already been inserted in the next available spot. The algorithm makes key comparisons,
continually exchanging keys until the heap property is satisfied or it reaches the root.
This pseudocode, as presented, assumes a tree-based implementation.

      Input : A heap H and an inserted node u
  1   curr ← u
  2   while curr.parent 6= φ and curr.key > curr.parent.key do
  3      swap curr.key, curr.parent.key
  4      curr ← curr.parent
  5   end

   Algorithm 28: Heapify: fixing a heap data structure after the insertion of a new
   node, u.
For the moment we will assume that we have an easy way to insert a node at the next
available spot. The heapify algorithm itself makes at most d comparisons and swaps
in the worst case (as was the case in the example in Figure 6.23). If we don’t need to



                                                                                     171
6 Trees




                                65                                                   65

                 32                       50                          32                       50

       25                  18        27        8            25                  90        27        8

  10        23        90                               10        23        18

(a) The new key, k = 90 is inserted at the next      (b) The first comparison finds that 90 and its
available spot.                                      parent 18 are out of order and exchanges them.

                                65                                                   90

                 90                       50                          65                       50

       25                  32        27        8            25                  32        27        8

  10        23        18                               10        23        18

(c) The second comparison ends up exchanging         (d) The third and final comparison ends up pro-
90 and 32.                                           moting 90 up to the new root.

                                Figure 6.23: Insertion and Heapification.




172
                                                                                   6.6 Heaps

                                                  50

                                        32               27

                                  25         18                8

                             10        23


                            Figure 6.24: Another Invalid Heap


heapify all the way up to the root node, then even fewer comparisons and swaps would
be necessary, but ultimately the process is O(d). However, since a heap guarantees
the fullness property, d = O(log n), the process only requires a logarithmic number of
comparisons/swaps in the worst case.


Retrieve Max

The other core operation is to get the maximum element. Because of the heap property,
this is guaranteed to be the root element. First, we remove the key/value from the
root node (which we have immediate access to) and save it off to eventually return it.
However, this leaves a gap in the tree that must be filled to preserve the fullness property.
Once again we may be tempted promote one of the root’s children to fill its place. To
preserve the heap property, we must promote the larger of the two children. Of course
we would have to continue down the heap as the promotion leaves another gap eventually
promoting a leaf element up. This idea alone will not suffice however. Again, consider
the heap from Figure 6.21. Were we to remove the root element 65 and proceed with the
operation as described we would move up 50 then 27 resulting in a “heap” as in Figure
6.24. Obviously this does not fulfill the fullness property as there is a gap (27’s left child
is missing).
As before, the solution is to prioritize the fullness property. Once we remove the root
element, we need to backfill it with a node that will preserve the fullness property,
specifically the “last” node (the rightmost node at the deepest level). Replacing the root
element with the last node may not necessarily satisfy the heap property, but we can
perform a top-down heapify process similar to what we did with the insert operation.
This process is illustrated properly in Figure 6.25.
We leave the development of the algorithm and pseudocode as an exercise (see Exercise
6.17). Assuming, as before, that we have easy access to the last element, then swapping
it and the the root is an O(1) operation. The heapification may again require up to d
comparisons and swaps if the last element must be exchanged all the way down to the
bottom of the heap. As before, since d = O(log n) this operation is O(log n).



                                                                                          173
6 Trees




                                                                              23

                  32                  50                         32                     50

       25              18        27          8             25            18        27          8

  10         23                                       10

(a) The root element, 65 is saved off in a tem-     (b) The last element, 23 replaces the root node.
porarily variable.                                  However, the heap property is not satisfied.

                            50                                                50

                  32                  23                         32                     27

       25              18        27          8             25            18        23          8

  10                                                  10

(c) The node is exchanged with the greater of the   (d) Another exchange must take place before the
two children to preserve the heap property.         heap is fixed.

            Figure 6.25: Removal of the root element (getMax) and Heapification




174
                                                                                6.6 Heaps

Secondary Operations

The heap properties don’t allow us to do general operations such as arbitrary search and
remove efficiently. We can achieve them using any of our tree traversal strategies, but all
of these operations have the potential to be O(n).
However, there are algorithms in which a heap is used that may require that keys in a
heap be changed (either increase or decrease their key values). A primary example is
when a heap is used to implement a priority queue and we wish to change the priority of
an element that is already enqueued.
First, we assume that we have “free” access to the node whose key we want to change as
finding an arbitrary node is going to be O(n) as previously noted. Once we have the
node u, we can increase or decrease its key value. Suppose we increase it. Its key value
was already greater than all of its descendants, so the heap property with respect to the
subtree rooted at u is still satisfied. However, we may have increased the key value such
it is now greater than its parent. We will need to once again, heapify and exchange keys
with its parents and ancestors until the heap property is once again satisfied. This is
exactly the same process as when we inserted a new node at the bottom.
Likewise, we can support a decrease key operation. Since the key value was already
smaller than its parent, the heap property is unaffected with respect to u’s ancestors.
In this case, however, the heap property could be violated with its children and/or
descendants. We simply exchange the key with the larger of its two children if the heap
property is violated and continue downward in the heap just as we did with the retrieve
(and remove) the maximum element. Both of.these operations are simply O(log n) as
before.


6.6.2 Implementations

So far we have presented heaps as binary trees. It is possible to implement heaps using the
same tree nodes and structure as with binary trees, but it does present some challenges
that we’ll deal with later. Another, easier and simpler implementation, however is to
come full circle back to array-based lists.


Array-Based Implementation

Recall that the fullness property of a heap essentially means that all nodes are contiguous.
It makes sense, then that we can store them in an array. In particular, we will store the
root element at index 1 (we will leave index 0 unused for this presentation though you
can easily shift all of the elements by one index if you’d rather not waste it).
Now, suppose a node u is stored at index i. Given this index, we need to locate u’s
parent and left and right child. These will each be stored at:



                                                                                       175
6 Trees

                                   i
                                    2
                                                               p

                              i
                                                      u
                                                                               2i + 1
                         2i
                                             `                 r

                         Figure 6.26: Heap Node’s Index Relations




                                                                         i
      0    1   2    3    4    5     6    7       8        9               2             i         2i 2i + 1

      –   65   32   50   25   18    27   8       10       23       ···    p    ···   u      ···   `    r




Figure 6.27: An array implementation of the heap from Figure 6.21 along with the
             generalized parent, left, and right child relations.




   • The left child is at index 2i
   • The right child is at index 2i + 1
   • The parent is at index 2i
                              

These relations are illustrated in Figure 6.26.
A full example is presented in Figure 6.27 which implements the same small max-heap
example we’ve been using. It also demonstrates the backward and forward relationships
to the parent and children nodes. Observe that the order in the array matches the order
from the tree if we were to perform a breadth first search on it (recall Figure 6.14).
A clear advantage to this implementation is that it is extremely easy and we can even
reuse an array-based list implementation. It is a simple matter to implement the heapify
algorithms in terms of indices given the mappings above (see Exercise 6.18). One
disadvantage is that it is costly to increase the size of the array when we need to add
more elements. However, amortized over the life of the heap and given the “savings” in
a simpler implementation, this may not be that big of a deal.



176
                                                                                  6.6 Heaps

Tree-Based Implementation

Though not common, you can implement a heap using the same binary tree structure
using a node with references to a parent, left child, and right child. However, to ensure
efficient operation for the heapify algorithms, we necessarily have to keep track of a
parent element for every node.
One problem is that we do not have obvious access to the “last” element or next available
spot in the heap as we did with an array-based implementation. With an array, we could
easily keep track of how many elements have been stored in the heap, say n. Then the
last element is necessarily at index n and the next available spot is at index n + 1. Since
we have random access, we can easily jump to either location to get the last element or
to insert at the next available spot.
With a tree structure, however, we do not have such access. Though we keep track of the
root, it is not straightforward to also keep track of the last element (or the next available
spot). Searching for either using BFS or another tree traversal algorithm would kill our
O(log n) efficiency.
Fortunately, using a bit of mathematical analysis and exploiting the fullness property of
a heap will allow us to find either the last element or the first available spot in O(log n)
time. We’ll focus on finding the first available open spot as the same technique can be
used to find the last element with minor modifications.
Without loss of generality, we’ll assume that we’ve kept track of the number of nodes in
the heap, n and thus we can compute the depth,

                                        d = blog nc

Due to the fullness property, simply knowing n and d give us a great deal of information.
In particular, we know that there are
                                      d−1
                                      X
                                            2k = 2d − 1
                                      k=0

nodes in the first d levels (levels 0 up to d − 1). Computing a simple difference,

                                     m = n − (2d − 1)

tells us how many nodes are in the last (deepest) level, level d. We also know that if the
last level were full, it would have 2d nodes. This tell us whether or not the next available
spot is in the left subtree or the right subtree by making a simple comparison:
               d
   • If m < 22 = 2d−1 then the left subtree is not “full” at level d and so it contains the
     next available spot.
   • Otherwise if m ≥ 2d−1 then the left subtree is full and the right subtree must
     contain the next available spot.



                                                                                         177
6 Trees

                                   r

                                                           d−1
                                                           X
                                                                 2k = 2d − 1
                                                           k=0

                      L                        R                  Number of
                                                                  nodes at level
                    level d                 level d               d is
                              2d                      2d
                                                                  m = n − (2d − 1)
                  at most      2
                                          at most      2



                       d                        d
              If m < 22 , the          If m ≥ 22 , the
              open slot is in          open slot is in
              the left subtree         the right subtree

Figure 6.28: Tree-based Heap Analysis. Because of the fullness property, we can determine
             which subtree (left or right) the “open” spot in a heap’s tree is by keeping
             track of the number of nodes, n. This can be inductively extended to each
             subtree until the open spot is found.




In each case we can traverse down to the left or right child respectively (which ever’s tree
contains the next available spot) and update both n and m (the number of elements at
level d) and repeat this process until we’ve found the next available spot. This analysis
is visualized in Figure 6.28 and the full process is presented as Algorithm 29.
It is not difficult to see that the complexity of this algorithm is O(d) = O(log n) since we
iterate down the depth of the heap using simple arithmetic operations. Thus the problem
of finding the last element or the next available spot does not significantly increase the
complexity of the heap’s core operations.



178
                                                                              6.6 Heaps


      Input : A tree-based heap H with n nodes
      Output : The node whose child is the next available open spot in the heap
  1   curr ← T.head
  2   d ← blog nc
  3   m←n
  4   while curr has both children do
  5      if m = 2d+1 − 1 then
            //remaining tree is full, traverse all the way left
  6         while curr has both children do
  7             curr ← curr.lef tChild
  8         end
  9      else
            //remaining tree is not full, determine if the next open
              spot is in the left or right sub-tree
                     d
 10         if m ≥ 22 then
               //left sub-tree is full
 11            d ← (d − 1)
                            d
 12             m ← (m − 22 )
 13             curr ← curr.rightChild
 14         else
               //left sub-tree is not full
 15            d ← (d − 1)
 16            m←m
 17            curr ← curr.lef tChild
 18         end
 19      end
 20   end
 21   output curr

          Algorithm 29: Find Next Open Spot - Numerical Technique



6.6.3 Variations

As presented in Definition 10, the heaps we have been describing are sometimes referred
to as binary heaps because they are based on a binary tree. There are other variations of
heaps that uses different structures and offer different properties.
For example, Binomial Heaps are heaps that are a collection of binomial trees. A



                                                                                     179
6 Trees

binomial tree of order k is a tree such that its children are binomial trees of order
k − 1, k − 2, . . . , 2, 1, 0 (that is, it has k children). This is an inductive definition so that
the base case of k = 0 is a single node tree. A binomial tree of order k has 2k nodes
and is of depth k. The core operations are a bit more complicated but have the same
complexity of O(log n). The advantage of this implementation is that a merge operation
of two heaps is also efficient (O(log n)).
A Fibonacci heap is a collection of trees that satisfy the heap property. However, the
structure is a lot more flexible than binary heaps or binomial heaps. The main advantage
is that some of the operations to keep track of elements are delayed until they are
necessary. Thus, some operations may be quite expensive, but when looked at from
an amortized analysis, the expected or average running time of the operations can be
interpreted as constant, O(1). This is similar to when we examined array-based lists.
Sometimes (though not often) we may need to expand the underlying array. Though
this is an expensive operation, since it is not too common, when you average the running
time of the core operations over the lifetime of the data structure, it all “evens out.” and
looks to be constant.


6.6.4 Applications

A heap is used in many algorithms as a data structure to efficiently hold elements to be
processed. For example, several graph algorithms such as Prim’s Minimum Spanning
Tree or Dijkstra’s Shortest Path algorithms use heaps as their fundamental building
block. A heap is also used in Huffman’s Coding algorithm to efficiently compress a file
without loss of information.


Priority Queue

From the core operations (insert and getMax) it might be obvious that a heap is an
efficient way to implement a priority queue. The enqueue operation is a simple insert and
a dequeue operation is always guaranteed to result in the maximum (highest priority)
element. Since both operations are O(log n), this is far more efficient than a list-based
priority queue. Note that the Java Collections library provides a heap-based priority
queue implementation in the java.util.PriorityQueue<E> class.


Heap Sort

Suppose we have a collection of elements. Now suppose we threw them all into a heap
and then, one-by-one, pulled them out using the getMax operation. As we pull them out,
what order would they be in? Sorted of course! By using a sophisticated data structure
like a heap, we can greatly simplify the code to achieve a more complex data operation,



180
                                                                                 6.6 Heaps

sorting. This is referred to as Heap Sort and is presented as Algorithm 30.

      Input : A collection of elements A = {a1 , . . . , an }
      Output : A, sorted
  1   H ← empty min-heap
  2   foreach a ∈ A do
  3      insert a into H
  4   end
  5   i←1
  6   while H is not empty do
  7      ai ← H.getM in
  8      i ← (i + 1)
  9   end
 10   output A

                             Algorithm 30: Heap Sort
Examine at the code in this algorithm. It is extremely simple; we put stuff into a
data structure, then we pull it out. It doesn’t get much simpler than that. The real
magic is in the data structure we used and exploited. This is a perfect illustration of
one of the themes of this book, borrowed from Eric S. Raymond [8] that “Smart data
structures and dumb code are a lot better than the other way around.” There are many
sophisticated sorting algorithms (Quick Sort, Merge Sort, etc.) that use clever code
(recursion, partitioning, etc.) and simple arrays. In contrast, Heap Sort uses very dumb
code: put stuff in, take stuff out and a very smart data structure. Beautiful.
Of course, this beauty is all for nothing if it is not also efficient. The complexity of the
insertion and getMin operations in Heap Sort actually changes on each iteration of the
algorithm because the data structure we’re using is growing and shrinking. Let’s first
analyze the “put stuff in” phase (the foreach loop, lines 2–3). On the first iteration, no
comparisons or swaps are made as H is initially empty. On the second iteration where
we insert the second element, H has size 1 and so 1 comparison (and potentially 1 swap)
is made. In general on the i-th iteration, H has i − 1 elements in it, thus the insertion
requires roughly log i − 1 (the depth of the heap) comparisons and/or swaps. This is all
summarized in Table 6.3.
Thus, the total number of comparisons (or swaps) made by Heap Sort is the summation
of the 4th column, or
                          n−1
                          X
                                log i = log 2 + log 3 + · · · + log n − 1
                          i=2

Note that we start the index at i = 2, since the first iteration does not require any com-
parisons. Using logarithm identities, we can further simplify and bound this summation:



                                                                                        181
6 Trees


                                Table 6.3: Analysis of Heap Sort
              Iteration      Size of H Depth of H Number of             Comparisons
                  1              0           –                          0
                  2              1           0                          1
                  3              2           1                          2
                  ..              ..         ..                         ..
                   .               .          .                          .
                  i            i−1        log (i − 1)               log (i − 1)
                  ..            ..              ..                        ..
                   .             .               .                         .
                  n           n−1         log (n − 1)               log (n − 1)



                       n−1
                       X
                             log i = log (2) + log (3) + · · · + log (n − 1)
                       i=2
                                  = log (2 · 3 · 4 · · · (n − 1))
                                  = log ((n − 1)!)
                                  ≤ log (n!)
                                  ≤ log (nn )
                                  = n log n

That is, the first phase is only O(n log n), matching the best case running time of Quick
Sort and other fast sorting algorithms. The analysis for the “take stuff out” phase is
similar, but in reverse. The size of the heap will be diminishing from n, n − 1, . . . , 1 and
will again contribute another factor of n log n. Since these are independent phases, the
entire algorithm is an efficient O(n log n) sorting algorithm. Simple, efficient, elegant and
beautiful.




182
                                                                          6.7 Exercises


6.7 Exercises

Exercise 6.1. Develop an algorithm to determine if a given tree T is a binary search
tree or not.
Exercise 6.2. Develop an algorithm to determine if a given tree T is a max heap or not.
Exercise 6.3. Adapt and implement the Stack-Based Preorder Traversal (Algorithm
18) for the Java BinaryTree<T> class.

Exercise 6.4. Adapt and implement the Stack-Based Inorder Traversal (Algorithm 21)
for the Java BinaryTree<T> class.

Exercise 6.5. Adapt and implement the Stack-Based Postorder Traversal (Algorithm
22) for the Java BinaryTree<T> class.

Exercise 6.6. Adapt and implement the Tree Walk Algorithm (Algorithm 24) for the
Java BinaryTree<T> class.
Exercise 6.7. Let T be a binary tree of depth d. What is the maximum number of
leaves that T could have with respect to d?
Exercise 6.8. Develop an algorithm that inserts a new node into a binary tree at the
shallowest available spot.
Exercise 6.9. Develop an algorithm that, given a binary tree T , counts the total number
of nodes in the tree (assume it is not kept track of as part of the data structure).
Exercise 6.10. Develop an algorithm that, given a binary tree T , counts the number of
leaves in it.
Exercise 6.11. Develop an algorithm that given a node u in a binary tree T determines
its depth.
Exercise 6.12. Develop an algorithm that given a binary tree T , computes its depth.
Exercise 6.13. Given an example of two trees with three nodes each that have a
different structure, but the same preorder traversal. Do the same for inorder and
postorder traversals.
Exercise 6.14. Write an algorithm that, given two binary trees, T1 , T2 , determines if
T1 = T2 . Two trees are equal if they have the same node structure and the same keys in
each node.
Exercise 6.15. Let T be a binary tree with n nodes. What is the maximum number of
leaves that T could have with respect to n? Provide an example for n = 15
Exercise 6.16. Develop an algorithm (write pseudocode) to insert a given key k into a
binary search tree T .



                                                                                    183
6 Trees

Exercise 6.17. Develop an algorithm (write pseudocode) fix a heap after its root element
has been removed.

Exercise 6.18. Rewrite Algorithm 28 (heapify) to work on an array instead of a tree
structure.

 Exercise 6.19. Design an algorithm to determine if two given binary trees T1 , T2 are
“equivalent” where the condition for equivalency is:
   1. The two trees contain the same set of keys.
   2. The two trees contain the same set of keys with the same structure.
   3. Repeat the first two conditions but when both T1 and T2 are binary search trees.




184
Appendix

1 Author-Book Database SQL

 1   drop table if exists AuthorBook;
 2   drop table if exists Book;
 3   drop table if exists Author;
 4

 5   create table Author (
 6     authorId int primary key auto_increment not null,
 7     firstName varchar(255) not null,
 8     lastName varchar(255) not null,
 9     key (lastName)
10   );
11

12   create table Book (
13     bookId int primary key auto_increment not null,
14     #isbn varchar(100) not null,
15     title varchar(255) not null,
16     numCopies int not null default 0,
17     key (title)
18     #unique key (isbn)
19   );
20

21   create table AuthorBook (
22     authorBookId int primary key auto_increment not null,
23     authorId int not null,
24     bookId int not null,
25     foreign key (authorId) references Author(authorId),
26     foreign key (bookId) references Book(bookId),
27     unique (authorId,bookId)
28   );
29

30   insert into Author (authorId, firstName, lastName) values
31     (1, "Norman", "Mailer"),
32     (2, "Douglas", "Adams"),



                                                                 185
Appendix


 33    (3, "Octavia", "Butler"),
 34    (4, "Cory", "Doctorow");
 35

 36   insert into Book (bookId, title, numCopies) values
 37     (1, "Naked and the Dead", 10),
 38     (2, "Dirk Gently's Holisitc Detective Agency", 4),
 39     (3, "The Hitchhiker's Guide to the Galaxy", 2),
 40     (4, "The Long Dark Tea-Time of the Soul", 1),
 41     (5, "Barbary Shore", 3),
 42     (6, "Kindred", 5);
 43

 44   insert into AuthorBook (authorBookId, authorId, bookId) values
 45     (1, 1, 1),
 46     (2, 2, 2),
 47     (3, 2, 3),
 48     (4, 2, 4),
 49     (5, 1, 5),
 50     (6, 3, 6);




186
Glossary
algorithm a process or method that consists of a specified step-by-step set of operations.

corner case a scenario that occurs outside of typical operating parameters or situations;
     an exceptional case or situation that may need to be dealt with in a unique or
     different manner. 66

data anomaly Redundant or inconsistent data in a database that violates the intended
     integrity of the data.. 5
dependency inversion The decoupling of software modules by defining an interface
    between them. Instead of one module depending directly on the other, both end
    up depending on the interface. One module implements the interface and the other
    (called a client) uses or consumes the interface. This allows client code to not
    have to depend on a particular implementation. Different implementations can be
    swapped out and changed with minimal code changes in the client code.. 53
duck typing In dynamically typed languages an object may not have a declared type
     and so to determine its type you rely on what method(s) and/or member variable(s)
     it has; “if it walks like a duck and talks like a duck” then it is a duck. That is, if it
     has the methods and variables that you expect of a particular type, then for all
     intents and purposes it is that type..

flat file A manner in which data is stored typically in a single file where the data model
       is “flattened” into a single table with many columns and each row representing a
       record.. 5

idiom in the context of software an idiom is a common code or design pattern. 62
interface segregation The principle that no client should be forced to depend on meth-
      ods or functionality it does not use. This principle essentially states that interfaces
      should be very minimal in their design and that many small interfaces should be
      combined to create more complex behavior..
iterator a pattern that allows a user to more easily and conveniently iterate over the
      elements in a collection. 64

leaky abstraction a design that exposes details and limitations of an implementation
     that should otherwise be hidden through encapsulation.. 73



                                                                                          187
Glossary

Liskov substitution principle The principle that if T is a subtype of S then any instance
     of S can be replaced with any subtype T without breaking a program..

mixin A mixin is a class that contains methods that other classes may use without a
     direct inheritance relationship. The functionality is “mixed in” to the class..

normalization the process of restructuring data and tables in a database, separating
    related data into their own tables in order to reduce redundancy and minimize the
    potential for data anomalies.. 47

open/closed principle The principle that “software entities should be open for extension,
    but closed for modification.” This generally refers to inheritance in OOP: that a
    class’s functionality can be extended through a subclass, but that the class itself
    should not be modifiable (so that other classes that depend on it don’t have the
    rug pulled out from under them)..

parameterized polymorphism a means in which you can make a piece of code (a method,
     class, or variable) generic so that its type can vary depending on the context in
     which the code is used. 63
polymorphism A mechanism by which a piece of code (class, method, variable) can be
     written generically so that it can “take on many forms” or used on different types
     in different places of a program..

query A request for data or an operation on data in a database. 7
queue a collection data structure that holds elements in a first-in first-out manner.

random access a mechanism by which elements in an array can be accessed by simply
     computing a memory offset of each element relative to the beginning of the array.
     68, 93

serialization translation of data into alternative formats, usually to plaintext data
      interchange formats such as JSON or XML. 6
single responsibility principle a general guideline that every module, class, function,
      etc. in code should have only a single responsibility or represent a single entity.
      More complex code can be written by composing these together..
stack a collection data structure that holds elements in a last-in first-out manner. 77
syntactic sugar syntax in a language or program that is not absolutely necessary (that
      is, the same thing can be achieved using other syntax), but may be shorter, more
      convenient, or easier to read/write. In general, such syntax makes the language
     “sweeter” for the humans reading and writing it. 64



188
                                                                                    Glossary

transaction The basic unit of work in a database that is treated as an atomic or an
     all-or-nothing event. A transaction may consist of one or more queries.. 7
tuple An ordered sequence of elements. Typically the notation (x1 , x2 , . . . , xn ) is used.
     Tuples correspond to rows or records in a database..




                                                                                         189
Acronyms
ACID Atomicity, Concurrency, Isolation, Durability. 7, 53
ADT Abstract Data Type. 57
API Application Programmer Interface.
AST Abstract Syntax Tree. 81

BFS Breadth First Search. 159
BLOB Binary Large Object. 13
BST Binary Search Tree. 161, 162

CRUD Create-Retrieve-Update-Destroy. 26, 37
CSV Comma-Separated Value. 5

DAG Directed Acyclic Graph.
DBA Database Administrator. 52
DDL Document Description Language. 9
DFS Depth First Search. 81, 147
DIP Dependency Inversion Principle.
DRY Don’t Repeat Yourself. 60

EDI Electronic Data Interchange. 6
ER Entity Relation. 20

FIFO First-In First-Out. 82, 83, 159
FK Foreign Key. 19
FOSS Free and Open Source Software. 9

GRASP General Responsibility Assignment Software Patterns. 3



                                                               191
Acronyms

ISP Interface Segregation Principle.

JDBC Java Database Connectivity API. 53
JPA Java Persistence API. 53
JSON JavaScript Object Notation. 6
JVM Java Virtual Machine. 63

KVP Key Value Pair. 53

LIFO Last-In First-Out. 77, 148
LSP Liskov Substitution Principle.

OCP Open-Closed Principle.
ORM Object-Relational Mapping. 53

PK Primary Key. 16

RDMS Relational Database Management System. 7
RTM Read The Manual. 13

SOLID SOLID Principles (Single Responsibility, Open/Closed, Liskov Substitution,
    Interface Segregation, Dependency Inversion.
SQL Structured Query Language. 9, 26
SRP Single Responsibility Principle.

XML Extensible Markup Language. 6




192
Index

aggregates, 37                          first normal form, 48
algorithm analysis, 105                 foreign key, 18
    recursive algorithms, 130           full outer join, 47
aliases, 31
amortized analysis, 129                 Gauss’s formula, 95
analysis of recursive algorithms, 130   group by clause, 38
array-based lists, 58                   heap sort, 180
asymptotic analysis, 108                heaps, 169
average case analysis, 128                 binomial heaps, 179
balanced binary search trees, 167          definition, 169
Big-O, 109                                 fibonacci heaps, 180
Big-Omega, 111                             heap sort, 180
Big-Theta, 112                           in operator, 34
binary search, 132                      inner join, see join
binary search tree, 161                 inorder traversal, 150
    definition, 161
binary search trees                     join
    balanced binary search trees, 167       cross join, 44
binary tree, 142                            full outer join, 47
binomial heaps, 179                         left join, 46
breadth first search, 159                join statement, 40
                                        join table, 23
cartesian product, 40
composite key, 24                       key
consumer producer pattern, 84                  composite, 24
cross join, 44                                 foreign, 18
                                               natural, 17
database, 5
                                               primary, 16
depth first search, 147
                                               surrogate, 17
deque, 87
                                               unique, 26
distinct clause, 33
                                        left join, 46
elementary operation, 107
                                         like clause, 33
entity-relation diagram, 20
                                        linked lists, 65
fibonacci heaps, 180                    lists, 57



                                                                  193
Index

      array lists, 58                    dropping, 15
      linked lists, 65               third normal form, 49
                                     tree, 137, 138
many-to-many, 22                         binary, 142
Master Theorem, 131                      binary search tree, 161
merge sort, 133                          definition, 139
                                         forest, 139
natural key, 17
                                         inorder traversal, 150
nested query, 29
                                         postorder traversal, 153
normal form
                                         preorder traversal, 148
    first normal form, 48
                                         traversal, 147
    second normal form, 48
    third normal form, 49            union, 47
normalization, 19, 47                union all, 47
nullability, 13                      unique key, 26
object oriented programming, 3       varchar, 12
one-to-many, 19
 order by clause, 35                 where clause, 32

postorder traversal, 153
preorder traversal, 148
primary key, 16
priority queue, 87
projections, 38
pseudocode, 102

queue
   priority queue, 87
queues, 82

relation, 41
relational database, 5

second normal form, 48
 select statement, 30
SQL, see Structured Query Language
stacks, 77
    peek, 80
Structured Query Language, 9, 26
surrogate key, 17

tables, 9
    altering, 15
    creation, 9
    data types, 11



194
Bibliography
[1] Manindra Agrawal, Neeraj Kayal, and Nitin Saxena. PRIMES is in P. Ann. of Math,
    2:781–793, 2002.
[2] P. Bachmann. Die analytische Zahlentheorie. Number v. 2 in Die analytische
    Zahlentheorie. Teubner, 1894.
[3] Arthur Cayley. On the theory of the analytical forms called trees. Philosophical
    Magazine Series 4, 13(85):172–176, 1857.
[4] Donald D. Chamberlin and Raymond F. Boyce. Sequel: A structured english query
    language. In Proceedings of the 1974 ACM SIGFIDET (Now SIGMOD) Workshop
    on Data Description, Access and Control, SIGFIDET ’74, pages 249–264, New York,
    NY, USA, 1974. ACM.
[5] E. F. Codd. A relational model of data for large shared data banks. Commun. ACM,
    13(6):377–387, June 1970.
[6] Leonhard Euler. Solutio problematis ad geometriam situs pertinentis. Commentarii
    academiae scientiarum Petropolitanae, 8:128–140, 1741.
[7] Donald E. Knuth. Big omicron and big omega and big theta. SIGACT News,
    8(2):18–24, April 1976.
[8] Eric S. Raymond. The Cathedral and the Bazaar. O’Reilly & Associates, Inc.,
    Sebastopol, CA, USA, 1st edition, 1999.




                                                                                195