The Engineering Method vs./and The Scientific Method

The Engineering Method: solving problems using heuristics (rules of thumb) that cause the best change in a poorly understood situation using available resources.

The Scientific Method: state a question, observe, state a hypothesis, test, analyze, and interpret

The public often confuses for the engineering method: In their minds, engineering is a subset of the scientific method, the public thinks of engineering as, to use a horrid term, “applied science.” The only place where the difference is clear is reflected in an old joke: “If it’s a success, then it’s a scientific miracle, if a disaster, then an engineering failure.”

The scientific and engineering methods have different goals: the scientific method wants to reveal truths about the universe, while the engineering method seeks solutions to real-world problems.

Bill Hammock 2021 speech for Hoover Medal: Reclaiming Engineering in the Minds of the Public – The Unheralded, Underappreciated, & Misunderstood Method that Built Our Modern World

One of Bill’s main points is that there is great creativity used in the everyday practice of engineering, and that story has largely been lost nowadays.

Unpacking Bill the Engineer Guy’s speech

Strongly recommend reading Bill’s entire speech linked here & above. The condensed story of the steam turbine below and/or video at the end of the page:

Charles Parsons – tinkerer and builder all his life. Worked on steam driven machinery for decades: “Under his father’s guidance, Parsons and his brother built, in 1869, a steam carriage that traveled seven miles per hour — a stunning device in an age where the horse remained supreme for several more decades.”

Henri Victor Regnault – largely forgotten French scientist. Meticulously collected data on the properties of steam. Engineers nowadays would these data the first steam tables.

Steam turbine – invented in 1885 by Charles Parsons by him building on the steam data of Regnault and literally thousands of years of human trying to build with steam power. Example: the aeolipile, designed by Hero of Alexandria in about 130 BC.

Parsons himself described the practical problems of executing his design as of “almost infinite complexity.”

What separated Parsons from the hundreds, perhaps, thousands of inventors before him, was how he navigated his way through this “infinite complexity.” Parsons was among the first engineers to be university-trained, similar to how we train engineers today. So, he turned to what he called the “data of the physicists.” In that data we see the role of science in engineering.

The patient and careful Regnault spent nearly thirty years documenting — as reported in these 3,000 pages — the thermodynamic properties of steam and other substances.

Regnault’s work helps dispel one myth about engineering and science: that a dramatic scientific breakthrough must precede a revolutionary new technology.

We see now the two things engineers need from science: high quality data and the some theory on how to calculate with that data. From the combined the work of Regnault and Rankine, Parsons knew, and I quote him here, that “a successful steam turbine ought to be capable of construction” because he could now size the number of wheels needed — thirty in his first successful turbine — and he could adjust the blades so that the steam flowed at the same rate through every section.

To say “science” created the turbine is to overlook Parsons’ great creativity, his superior machining, and the ten years of trial and error needed to refine the turbine. To call Parson’s work [engineering] “applied science” is the fuzziest of thinking: it conflates the tool with the method. It’s akin to saying that carpentry is “applied hammering,” that composing music is “applied pitch,” or that writing a book is “applied lettering.”

The solitary genius description of invention

Love that Bill takes aim at the Great Man Theory in this section of the speech:

Whenever we reduce any engineering achievement to one single cause — to the discovery of a scientific fact or even the development of the first working prototype — we hide the rich creativity of engineers from the public. For example, we’re all familiar with the story of Edison and the light bulb: once he discovered the proper filament — carbonized bamboo — the story ends. I know that we all love stories of sole inventors whose spark of inspiration revolutionized the world. They give us narratives that are neat, tidy, and digestible, but incomplete. It hides the engineering method; it conceals the creativity of engineers, smooths over struggles, and sanitizes choice that reflects cultural norms. A technology like a light bulb only solves problems when it can be manufactured or mass produced. A handful of working light bulbs in the late 1800s is a marvel, but it doesn’t light the world. In this sense, the “invention” of the incandescent light bulb is a decades-long process of incremental changes to create a filament that can be manufactured reliably. To tell only a “solitary genius” story shortchanges the contributions of inventive and imaginative engineers who were essential to a technology’s development.

For example, the creativity of Lewis Latimer who devised novel methods to reliably manufacture and assemble carbon filaments. His work was the industry’s standard for the first decade of the commercial light bulb — a critical period that cemented the light bulb as essential — until the carbon filament was replaced by ductile tungsten by William Coolidge — another untold story, yet also an exemplar of the engineering method.

Engineering Method x Cynefin

A “rule of thumb,” otherwise called, more formally, a “heuristic” is an imprecise method used as a shortcut to find the solution to a problem. It is an idea so old and pervasive that practically every language seems to have its own corresponding term, uncannily following a theme of body parts: in French “the nose”, in German “the fist,” in Japanese “measuring with the eye,” and in Russian “by the fingers.” All expressing an imprecise method of guidance by common knowledge, a protocol of estimation. In practice it’s anything that can plausibly aid the solution of a problem, but is not justified from a scientific or philosophical perspective either because it doesn’t need to be or simply because it can’t be justified through anything other than results. Rather than define, it’s best to list the four key characteristics of a rule of thumb.

These four characteristics can be illustrated with a simple rule of thumb used to improve a player’s chess game: “Control the center of the board.”

  1. A rule of thumb reduces the time and effort needed to search for a solution to a problem.
  2. It can secure a probability of success, but it does not guarantee success.
  3. It can remain valid while simultaneously contradicting other rules of thumb that help solve the same problem.
  4. It rejects absolute standards. Rules of thumb are designed to be applied and judged according to a problem’s context, but become less useful, perhaps even meaningless, when considered abstractly or objectively.

“Control the center of the board” is an excellent example of an enabling constraint – a good strategy when navigating a Complex domain

Engineering Method x DRRS

In DRRS MOOC #1 module 1.2-5-4, the article on Regenerative Agriculture (https://www.nzgeo.com/stories/regeneration/) had a recurring theme about “does the science support (or prove) regenerative agriculture is better?”
The answer was sometimes yes, sometimes no, sometimes maybe/inconclusive – and author does a nice job pointing out the limits of scientific inquiry in this area and the benefits beyond crop yields that the scope of science often does not measure.

To me, the Engineering Method sounds a lot like TEK (Traditional Ecological Knowledge / Indigenous Knowledge): heuristics are a living body of knowledge, often steeped in a time and a place, passed down generation to generation.


I consider myself an engineer and a scientist, but I think the “sciencification” of everything can be become a hindrance – it’s disempowering to humans everywhere to say only act if there’s definitive, science-proven evidence. While we should be concerned that we get good outcomes from the time & energy invest, we don’t necessarily need to know every lever, every causal link before taking action. As Bill the Engineer Guy points out, engineering (or designing) often comes first, and science figures out why it works after (if the scientific study can find funding!).

Additional videos in Bill’s series