previous chapter
Design of Algorithms
next chapter

Design of Algorithms

C. L. Liu

C. L. Liu obtained a bachelor of science degree in 1956 at Cheng Kung University in Tainan, Republic of China, Taiwan. In 1960 and 1962, respectively, he earned a master's degree and a doctorate in electrical engineering at MIT. Currently, he is Professor of Computer Science at the University of Illinois-Urbana/Champaign. Dr. Liu's principal research interests center on the design and analysis of algorithms, computer-aided design of integrated circuits, and combinatorial mathematics.

Let me begin by telling you a story I heard. There was an engineer, a physicist, and a mathematician. They all went to a conference together.

In the middle of the night, the engineer woke up in his room and discovered there was a fire. He rushed into the bathroom, got buckets and buckets of water, threw them over the fire, and the fire was extinguished. The engineer went back to sleep.

In the meantime, the physicist woke up and discovered there was a fire in his room. He jumped out of bed, opened his briefcase, took out paper, pencil, and slide rule, did some very careful computations, ran into the bathroom, measured exactly one-third of a cup of water, and poured it slowly over the fire. Of course, the fire was extinguished. And the physicist went back to sleep.

In the meantime, the mathematician woke up and also discovered there was a fire in his room. He jumped out of bed, rushed into the


260

bathroom, turned on the faucet—and he saw water come out. He said, "Yes, the problem has a solution." And he went back to sleep.

Today we are here to talk about design of algorithms. Indeed, we sometimes use the engineer's approach, and sometimes the physicist's approach, and occasionally the mathematician's approach.

Indeed, it is a fair question that many of you can stare into the eyes of the algorithmists and say that throughout the years we have given you faster and faster and cheaper and cheaper supercomputers, more and more flexible software. What have you done in terms of algorithmic research?

Unfortunately, we are not in a position to tell you that we now have a computer capable of 1012 floating-point operations per second (TFLOPS) and that, therefore, there's no need to do any more research in the algorithms area. On the other hand, we cannot tell you that our understanding of algorithms has reached such a point that a "uniFLOPS" computer will solve all the problems for us. What is the problem? The problem is indeed the curse of combinatorial explosion.

I remember it was almost 30 years ago when I took my first course in combinatorics from Professor Gian-Carlo Rota at MIT, and he threw out effortlessly formulas such as n5 , n log n, 2n , 3n , and nn . As it turns out, these innocent-looking formulas do have some significant differences. When we measure whether an algorithm is efficient or not, we draw a line. We say an algorithm is an efficient algorithm if its computational time as a function of the size of the problem is a polynomial function. On the other hand, we say that an algorithm is inefficient if its computational time as a function of the size of the problem is an exponential function. The reason is obvious. As n increases, a polynomial function does not grow very fast. Yet on the other hand, an exponential function grows extremely rapidly.

Let me just use one example to illustrate the point. Suppose I have five different algorithms with these different complexity measures, and suppose n is equal to 60. Even if I'm given a computer that can carry out 109 operations per second, if your algorithm has a complexity of n, there is no problem in terms of computation time (6 × 10-8 seconds). With a complexity of n2 or n5 , there is still no problem. There is a "small" problem with a complexity of 2n and a "big" problem with a complexity of 3n .

Now, of course, when the complexity is 3n , computation time will be measured in terms of centuries. And measured in terms of cents, that adds up to a lot of dollars you cannot afford. As Professor Rota once told us, when combinatorial explosion takes place, the units don't make any difference whatsoever.


261

So indeed, when we design algorithms it is clear that we should strive for efficient algorithms. Consequently, if we are given a problem, it would be nice if we can always come up with efficient algorithms.

On the other hand, that is not the case. A problem is said to be intractable if there is known to be no efficient algorithm for solving that problem.

From the algorithmic point of view, we have paid much less attention to problems that are really intractable. Rather, a great deal of research effort has gone into the study of a class of problems that are referred to as NP-complete problems. We have tackled them for 40 or 50 years, and we cannot confirm one way or the other that there are or are not efficient algorithms for solving these problems.

So therefore, although I begin by calling this the curse of combinatorial explosion, it really is the case, as researchers, that we should look at this as the blessing of combinatorial explosion. Otherwise, we will have no fun, and we will be out of business.

So now the question is, if that is the case, what have you been doing?

Well, for the past several years, people in the area of algorithmic design tried to understand some of the fundamental issues, tried to solve real-life problems whose sizes are not big enough to give us enough trouble, or, if everything failed, tried to use approximation algorithms that would give us good, but not necessarily the best possible, results.

Let me use just a few examples to illustrate these three points.

I believe the problem of linear programming is one of the most beautiful stories one can tell about the development of algorithms.

The problem of linear programming was recognized as an important problem back during the Second World War. It was in 1947 when George Dantzig invented the simplex method. Conceptually, it is a beautiful, beautiful algorithm. In practice, people have been using it to solve linear programming problems for a long, long time. And indeed, it can handle problems of fair size; problems with thousands of variables can be handled quite nicely by the simplex method.

Unfortunately, from a theoretical point of view, it is an exponential algorithm. In other words, although we can handle problems with thousands of variables, the chance of being able to solve problems with hundreds of thousands of variables or hundreds of millions of variables is very small.

In 1979, Khachin discovered a new algorithm, and that is known as the ellipsoid algorithm. The most important feature of the ellipsoid algorithm is that it is a polynomial algorithm. The algorithm by itself is


262

impractical because of the numerical accuracy it requires; the time it takes to run large problems will be longer than with the simplex method.

But on the other hand, because of such a discovery, other researchers got into the picture. Now they realize the ball game has moved from one arena to a different one, from the arena of looking for good exponential algorithms to the arena of looking for good polynomial algorithms.

And that, indeed, led to the birth of Karmarkar's algorithm, which he discovered in 1983. His algorithm is capable of solving hundreds of thousands of variables, and, moreover, his research has led to many other activities, such as how to design special-purpose processors, how to talk about new previous hit architectures next hit, and how to talk about new numerical techniques. As I said, this is a good example illustrating algorithmic development. It demonstrates why it is important to have some basic understanding of various features of algorithms.

Let me talk about a second example, with which I will illustrate the following principle: when it becomes impossible to solve a problem precisely—to get the exact solution—then use a lot of FLOPS and try to get a good solution. And I want to use the example of simulated annealing to illustrate the point.

Back in 1955, Nick Metropolis, of Los Alamos National Laboratory, wrote a paper in which he proposed a mathematical formulation to model the annealing process of physical systems with a large number of particles. In other words, when you have a physical system with a large number of particles, you heat the system up to a high temperature and then slowly reduce the temperature of the system. That is referred to as annealing. Then, when the system freezes, the system will be in a state of minimum total energy. In Nick's paper he had a nice mathematical formulation describing the process.

That paper was rediscovered in 1983 by Scott Kirkpatrick, who is also a physicist. He observed that the process of annealing is very similar to the process of doing combinatorial minimization because, after all, you are given a solution space corresponding to all the possible configurations that the particles in the physical system can assume. If, somehow, you can move the configurations around so that you can reach the minimum energy state, you would have discovered the minimum point in your solution space. And that, indeed, is the global minimum of your combinatorial optimization problem.

The most important point of the annealing process is, when you reduce the temperature of your physical system, the energy of the system does not go down all the time. Rather, it goes down most of the time, but it goes up a little bit, and then it goes down, and it goes up a little bit again.


263

Now, in the terminology of searching a large solution space for a global minimum, that is a very reasonable strategy. In other words, most of the time you want to make a downhill move so that you can get closer and closer to the global minimum. But occasionally, to make sure that you will not get stuck in a local minimum, you need to go up a little bit so that you can jump out of this local minimum.

So therefore, as I said, to take what the physicists have developed and then use that to solve combinatorial optimization problems is conceptually extremely pleasing.

Moreover, from a mathematical point of view, Metropolis was able to prove that as T approaches infinity, the probability that the system will reach the ground state approaches 1 as a limit. For many practical problems that we want to solve, we cannot quite make that kind of assumption. But on the other hand, such a mathematical result does give us a lot of confidence in borrowing what the physicists have done and using that to solve optimization problems.

As I said, the formula I have here tells you the essence of the simulated-annealing approach to combinatorial optimization problems. If you are looking at a solution S and you look at a neighboring solution S', the question is, should you accept S' as your new solution?

This, indeed, is a step-by-step sequential search. And according to the theory of annealing, if the energy of the new state is less than the energy of the current state, the probability of going there is equal to one.

On the other hand, if the new state has an energy that is larger than the energy of the current state, then it depends on some probabilistic distribution. What I'm trying to say is, what I have here is nothing but some kind of probabilistic uphill/downhill search. And as it turns out, it will go quite well in many combinatorial-optimization problems.

And moreover, the development of such techniques will lead to a lot of interesting research questions. How about some theoretical understanding of the situation? How about possible implementation of all of these ideas? And moreover, simulated annealing basically is a very sequential search algorithm. You have to look at one solution, after that the next solution, after that the next solution. How do you parallelize these algorithms?

Let me quickly mention a rather successful experience in solving a problem from the computer-aided design of integrated circuits. That is the problem of placement of standard cells. What you are given is something very simple. You are given a large number, say about 20,000 cells. All these cells are the same height but variable widths. The problem is how to place them in rows. I do not know how many rows all together,


264

and I do not know the relationship among these cells except that eventually there will be connections among them. And if you do it by brute force, you are talking about 20,000! possibilities, which is a lot.

Yet, on the other hand, there is a software package running on Sun Microsystems, Inc., workstations for about 10 hours that produces solutions to this problem that are quite acceptable from a practical point of view. That is my second example.

Let me talk about a third example. As I said, since we do not have enough time, enough FLOPS, to get exact solutions of many of the important problems, how about accepting what we call approximate solutions? And that is the idea of approximation algorithms.

What is an approximation algorithm? The answer is, engineers wave their hands and yell, "Just do something!"

Here is an example illustrating that some basic understanding of some of the fundamental issues would help us go a long way. Let me talk about the job-scheduling problem, since everybody here who is interested in parallel computation would be curious about that. You are given a set of jobs, and these jobs are to be scheduled—in this example, using three identical processors. And of course, when you execute these jobs, you must follow the precedence constraints.

Question: how can one come up with a schedule so that the total execution time is minimum? As it turns out, that is a problem we do not know how to do in the sense that we do not know of any polynomial algorithm for solving the problem. Therefore, if that is the case, once you have proved that the problem is NP-complete, you are given a license to use hand-waving approximation solutions.

However, if you are given a particular heuristic, a particular hand-waving approach, the question is, what is the quality of the results produced by your particular heuristics? You can say, I will run 10,000 examples and see what the result looks like. Even if you run 10,000 examples, if you do not know the best possible solution for each of these examples, you're getting nowhere.

In practice, if I give you just one particular instance and I want you to run your heuristics and tell me how good is the result that your heuristic produces, how do you answer the question? As it turns out, for this particular example, there is a rather nice answer to that question. First of all, let me mention a very obvious, very trivial heuristic we all have been using before: whenever you have a processor that is free, you will assign to it a job that can be executed.

In this case, at the very beginning, I have three processors. They are all free. Since I have only one job that can be executed, I'll execute it. After


265

that, I have three processors that are free, but I have four jobs that can be executed. I close my eyes and make a choice; it's B, C, E. Do the rest, and so on and so forth.

In other words, the simple principle is, whenever you have jobs to be executed, execute them. Whenever your processors are free, try to do something on the processors. This is a heuristic that is not so good but not so bad, either. It can be proved that if you follow such a heuristic, then the total execution time you are going to have is never more than 1.66 times the optimal execution time. In the worst case, I will be off by 66 per cent. Now, of course, you told me in many cases 66 per cent is too much. But there's another way to make good use of this result.

Suppose you have another heuristic which is, indeed, a much better heuristic, and your total execution time is 'instead of  . How good is'? Ah, I will compare it with my , since it is very easy for me to compute . If your ' is close to the value of , you might be off by 66 per cent.

On the other hand, if your ' is close to 3/5 of , you are in pretty good shape because you must be very close to  0 . This is an example to show you that although we do not know how to find  0 , although we do not know how to determine a schedule that will give me an optimal schedule, on the other hand, we will be capable of estimating how good or how bad a particular heuristic is, using it on a particular example.

To conclude, let me ask the question, what is the research agenda? What are we going to do?

First point: we should pay more attention to fundamental research. And it has been said over and over again, theoretical research has been underfunded. As a college professor, I feel it is alarming that our Ph.D. students who have studied theoretical computer science very hard and have done very well face very stiff competition in getting jobs in good places. There is a danger some of our brightest, most capable computer-science students would not want to go into more theoretical, more fundamental, research but would rather spend their time doing research that is just as exciting, just as important, but fosters better job opportunities.

I think we should look into possibilities such as arranging postdoctoral appointments so that some of the people who have a strong background in theory can move over to learn something about other areas in computer science while making use of what they learned before in theoretical computer science.

Second point: with the development of fast supercomputers, there are many, many opportunities for the theoreticians to make use of what they learn about algorithms, to solve some real live application problems.


266

Computer-aided design is an example; image-processing, graphics, and so on are all examples illustrating that theoreticians should be encouraged to make good use of the hardware and the software that are available so that they can try to help to solve some of the important problems in computer science.

Third point: when we compare combinatorial algorithms with numerical algorithms, combinatorial algorithms are way behind in terms of their relationship with supercomputing, and we should start building a tool library—an algorithm library—so that we can let people who do not know as much, who do not care to know as much, about algorithms make good use of what we have discovered and what we have developed.

And finally, I'm not saying that as theoreticians we necessarily need TFLOPS machines, but at least give us a lot more FLOPS. Not only will we make good use of the FLOPS you can provide, but you can be sure that we'll come back and beg you for more.


267

previous chapter
Design of Algorithms
next chapter