Experiments on Enzyme Activity | Biochemistry

experiment about enzymes

ADVERTISEMENTS:

The below mentioned article includes a collection of seven experiments on enzyme activity.

1. Experiment to demonstrate the activity of enzymes:

Requirements:

Benzidine solution, razor, thin sections of actively growing root (or germinating seeds or germinating pollen grains), phosphate buffer, hydrogen peroxide (1%), ammonium chloride (5%), starch paste.

Method, observations and results:

1. Cut thin sections of fast-growing roots or germinating seeds or take some pollen grains and put the material in phosphate buffer at pH 7.0. Now transfer the material in the incubation mixture [made of 5% benzidine saturated solution + 5 ml hydrogen peroxide (1%) and 1 ml ammonium chloride (5%)]. Incubate the entire mixture at room temperature for 5 minutes and observe. Dark-blue colour appears. This is due to the presence of the enzyme peroxidase.

2. Make a paste of starch, immerse in it the provided material like actively growing roots or germinating pollen grains and test for sugar. Positive test of sugar indicates that starch has transformed into sugars, and this further confirms the presence of enzyme amylase in the studied plant material.

2. Experiment to demonstrate that heat destroys activity of enzyme but not that of a catalyst:

Four test tubes, manganese dioxide (MnO 2 ), water, beaker, spirit lamp, piece of fresh liver or potato, hydrogen peroxide, boiled and cooled piece of liver or potato, boiled and cooled manganese dioxide solution.

Pour 2 ml of H 2 O 2 solution is each of the four test tubes. Add a pinch of manga­nesedioxde in the first test tube, pre-boiled and cooled 1 ml of manganese dioxide solution in the second, small piece of fresh liver or potato in third and pre-boiled and cooled piece of liver or potato in the fourth.

All the test tubes are kept at room temperature if it is summer and warm water (kept at about 38°C) if it is winter. Oxygen bubbles are found to come out of solution in the first three test tubes but not in the fourth one.

Manganese dioxide is a catalyst that helps to break hydrogen peroxide into water and oxygen. The catalyst is not affected by heat as it is clear from evolution of oxygen in the second test tube.

Fresh liver or potato contains enzymes peroxidase and catalase that help in evolving oxygen from hydrogen peroxide. The enzymes are functional at room temperature (test tube three) but heating destroys their activity (test tube four). Therefore, heat kills enzyme activity but has little effect on the activity of a catalyst.

Precautions:

(i) Test tubes must be thoroughly washed with water before use.

(ii) Do not perform the experiment without warm water if it is cold outside,

(iii) Keep the temperature of warm water to about 38″C. Do not allow it to go beyond 50°C.

(iv) Use fresh potato slice or liver piece.

3. Experiment to prove that enzymes are specific in their activity:

Four test tubes, one percent starch solution, one percent sucrose solution, saliva, Bennedict’s or Fehling’s solution, spirit lamp.

Take four test tubes. Pour 1 ml of 1% starch solution in the first, 1 ml of 1% sucrose solution in the second, 1 ml each of 1% starch solution and saliva in the third, and 1 ml each of 1% sucrose solution and saliva in the fourth. Keep all the test tubes undisturbed for one hour at room temperature if it is summer and in warm water (about 38°C) if winter.

Afterwards pour 3 ml of Bennedict’s or Fehling’s solution (having cupric ions) in each of the test tube. Heat the test tube to boiling for about 2 minutes and observe. There is no change of colour in first and second test tubes. The fourth test tube also does not show any change in colour. However, in the third test tube, the blue colour of Bennedict’s or Fehling’s solution gets changed to yellowish or reddish precipitate.

The change of colour from blue to yellowish or reddish precipitate is caused by conversion of cupric ions of Bennedict’s or Fehling’s solution into cuprous oxide.

This occurs in the presence of reducing sugars. Both sucrose and starch are non-reducing in nature as is found by the absence of colour change in test tube one and two. In test tube four having sucrose and saliva there is no enzyme activity since colour change is absent.

However, test tube three having starch and saliva (containing enzyme salivary amylase) reducing sugars are produced as is clear from change of colour. Therefore, the enzyme, salivary amylase, present in saliva catalyses hydrolysis of starch but not sucrose which is a common disaccha­ride.

(i) Test tubes must be thoroughly washed and dried;

(ii) The water in which the solutions are to be kept must not be very hot.

(iii) In winter warm water must be taken,

(iv) Before the collection of saliva, mouth must be washed thoroughly with distilled water,

(v) Swill the rinsed mouth with 7-10 cc of distilled water for 1 minute and then collect the same, as water containing saliva,

(vi) While heating the solution should not be allowed to bump violently.

4. Experiment to demonstrate the activity of the enzyme amylase extracted from the germinating barley or pea seeds:

Starch powder, iodine solution, germinating barley or pea seeds, distilled water, test tubes, mortar, pestle, filter paper, and funnel.

1. Take a little amount of ordinary starch and make a thin paste of it in about 50 ml boiling water. Allow it to cool down.

2. Take about 5 gm. of germinating seeds of barely or pea. In case of pea seedlings, remove their cotyledons. Ground the cotyledons along with distilled water in the mortar, and filter the contents through a funnel.

3. Pour the starch solution in two test tubes and mark them as ‘A’ and ‘B’.

4. Add a few drops of iodine solution in tube ‘A’ and observe the colour change.

5. Add the filterate of the crushed cotyledons or endosperm of barely in tube ‘B’.

6. Keep both the tubes at a warm place (about 35°- 40°C) for about 30 minutes.

Observations:

In test tube ‘A’, the contents turn blue in colour. In test tube ‘B’, the contents show reddish- brown colour. After about 15 minutes, if iodine solution is added, it shows no positive test for starch in tube ‘B’. However, if a few drops of Fehling’s solution are added in tube ‘B’, a brick-red precipitate appears.

Formation of blue colour in tube ‘A’ confirms the test of starch. In tube ‘B’, formation of reddish-brown colour is due to the fact that addition of seed extract supplied the enzyme amylase which partially hydrolysed the starch into maltose (a 12-carbon sugar). After about 30 minutes, the entire starch in tube ‘B’ gets completely hydrolysed into hexose sugar.

Due to this the iodine solution gives negative test for starch. Formation of reddish-brown colour after the addition of Fehling’s solution confirms the presence of hexose sugar. It is a reducing sugar, and a product of hydrolysis of starch made of amylase and amylopectin.

The enzyme amylase is present in the germinating barley or pea seeds. It is released during the crushing process. Amylase is actually an enzyme which catalyzes the breakdown of starch into monosaccharide units.

5. Experiment to study the enzyme activity of diastase in germinating seeds of barley and to study the influence of pH and temperature:

The enzyme diastase acts on starch and converts it to hexose sugar.

Germinating barley seeds, mortar, water, muslin cloth, centrifuge, measuring flask, iodine solution prepared in potassium iodide), starch solution, test tubes, beaker, enzyme extract, pipette, Benedict’s solution, buffer solutions of known pH, diastase solution, water baths (7), stop watch.

Preparation of Required Solutions:

(a) Starch solution:

It is prepared by adding 2 gm. of soluble starch in 50 ml of boiling water.

(b) Buffer solution A:

Add 6.95 gm. of monobasic sodium phosphate (0.2 M) in 250 ml of distilled water and use it as buffer solution A.

(c) Buffer solution B:

Dissolve 17.92 mg of dibasic sodium phosphate (Na 2 HPO 4 .12H 2 O) in 250 ml of water to get 0.2 M buffer solution B.

(d) Iodine solution (1 %):

Mix 1 gm. iodine and 2 gm. KI in 300 ml water and use it as iodine solution.

Method and observations:

1. Take 10 gm. of germinating barley seeds and 20 ml of water and grind them in mortar.

2. Filter the above mixture through muslin cloth, centrifuge the filtrate at low speed, take the supernatant liquid in a measuring flask and make up the volume of the enzyme extract upto 100 ml.

3. Take six test tubes and put 1 ml iodine solution (1 %) in each of them. Also add 20 ml water in each of them.

4. In a separate test tube take 1 ml of 1% iodine solution and 20 ml water and add 1 ml starch solution. This will work as a control.

5. Take 10 ml of starch solution in a beaker, add 1ml enzyme extract and shake it well.

6. This starch diastase mixture is now called digestion mixture. Pipette out 1ml of this digestion mixture and add it in each of the six test tubes containing iodine solution and observe. Colour starts disappearing. Note the time of disappearance of colour.

7. After about 10 minutes of digestion, take out 1ml of digestion mixture into a test tube, and now test for sugar by Benedict’s reagent. Sugar test is positive.

This shows that activity of the enzyme diastase has transformed the starch into sugar. This further confirms the enzyme activity of diastase in germinating seeds of barley.

Effect of pH:

Take 9 test tubes and in each of them take 5ml of buffer solution of known pH like 5.0,5.5, 6.0,6.5,7.0,7.5,8.0,8.5,9.0. In each test tube, add 5 ml of starch solution (1%). Now add 1 ml of diastase solution in each of the test tubes and note the time of addition. Shake the contents thoroughly and mix them well.

Now pipette 1 ml of the reaction mixture every 5 minutes into separate test tubes containing 1ml of iodine solution and 20 ml of water. Take a graph paper and plot the time taken in minutes for complete hydrolysis (as shown by complete disappearance of colour of iodine) against the pH. Time of disappearance of colour is different in different pH, and this shows the influence of pH on enzyme activity.

Effect of temperature:

Take seven water baths and maintain them at seven different temperatures like 100°C, 80°C, 60°C, 40°C, 20°C, 10°C and 0°C. Take seven test tubes and in each of them add 5 ml of soluble starch (1 %) maintained at pH 7 and put them in seven different water baths maintained at different temperatures.

Note the time when contents in different test tubes attain the temperature of their respective water baths and now add 1 ml of diastase solution in each test tube. Shake the test tubes well, wait for 5 minutes and add in each test tube 1 ml mixture of 1 ml iodine solution and 20 ml water. Wait for some time, note the time taken for disappearance of iodine colour against each temperature and plot the time on graph paper.

6. Experiment to demonstrate the activity of peroxidase in plant material:

Potato tuber, test tube, muslin cloth, hydrogen peroxide (3%), pyrogallol solution (1%).

1. Macerate 5 gm. of potato tuber, squeeze it through a muslin cloth, and take 3 ml of the potato extract in a test tube.

2. Put 1% solution of pyrogallol (a phenolic compound) in the test tube containing potato extracts and add 3 drops of hydrogen peroxide (3%) and observe.

Observation:

Change in colour takes place.

This change in colour shows the presence of peroxidase activity in the extract of potato tuber.

7. Experiment to demonstrate the pH change inhibits in enzyme activity:

Two test tubes. 1% starch solution, saliva, dilute HCI, beaker, water, spirit lamp, iodine solution (I + KI).

Pour 2 ml of starch solution in each of the two test tubes. Add 1 ml of fresh saliva in each. A few drops of dilute hydrochloric acid are added to one of the test tubes to make its solution acidic. Both the test tubes are kept for one hour at room temperature if it is summer or in warm water (about 38°C) if it is winter.

After one hour both the test tubes are tested for starch by pouring 1 to 2 drops of iodine solution. Test tube one shows negative starch test while test tube two (acidified) develops blue colour showing the presence of starch.

In test tube one, pH is nearly that of saliva. The salivary amylase contained in saliva is functional and causes hydrolysis of starch because starch test is negative.

In test tube two appearance of blue colour indicates that the starch has not been hydrolysed by enzyme present in saliva. The only difference in two test tubes is that the solution of the second test tube has been acidified. Therefore, change in pH inhibits enzyme activity.

(i) Test tubes must be thoroughly washed and dried,

(ii) Before the collection of saliva mouth must be washed and any acidic or alkaline food must not be taken,

(iii) Warm water must be used in winter,

(iv) Care must be taken that there is no contamination of first test tube by the HCI which is being used for second test tube.

Related Articles:

  • Detection of Enzymatic Activity | Plant Tissue
  • Estimation of Amylase | Enzyme Activity

Experiment , Biochemistry , Enzymes , Experiments on Enzyme Activity

  • Anybody can ask a question
  • Anybody can answer
  • The best answers are voted up and rise to the top

Forum Categories

  • Animal Kingdom
  • Biodiversity
  • Biological Classification
  • Biology An Introduction 11
  • Biology An Introduction
  • Biology in Human Welfare 175
  • Biomolecules
  • Biotechnology 43
  • Body Fluids and Circulation
  • Breathing and Exchange of Gases
  • Cell- Structure and Function
  • Chemical Coordination
  • Digestion and Absorption
  • Diversity in the Living World 125
  • Environmental Issues
  • Excretory System
  • Flowering Plants
  • Food Production
  • Genetics and Evolution 110
  • Human Health and Diseases
  • Human Physiology 242
  • Human Reproduction
  • Immune System
  • Living World
  • Locomotion and Movement
  • Microbes in Human Welfare
  • Mineral Nutrition
  • Molecualr Basis of Inheritance
  • Neural Coordination
  • Organisms and Population
  • Photosynthesis
  • Plant Growth and Development
  • Plant Kingdom
  • Plant Physiology 261
  • Principles and Processes
  • Principles of Inheritance and Variation
  • Reproduction 245
  • Reproduction in Animals
  • Reproduction in Flowering Plants
  • Reproduction in Organisms
  • Reproductive Health
  • Respiration
  • Structural Organisation in Animals
  • Transport in Plants
  • Trending 14

Privacy Overview

CookieDurationDescription
cookielawinfo-checkbox-analytics11 monthsThis cookie is set by GDPR Cookie Consent plugin. The cookie is used to store the user consent for the cookies in the category "Analytics".
cookielawinfo-checkbox-functional11 monthsThe cookie is set by GDPR cookie consent to record the user consent for the cookies in the category "Functional".
cookielawinfo-checkbox-necessary11 monthsThis cookie is set by GDPR Cookie Consent plugin. The cookies is used to store the user consent for the cookies in the category "Necessary".
cookielawinfo-checkbox-others11 monthsThis cookie is set by GDPR Cookie Consent plugin. The cookie is used to store the user consent for the cookies in the category "Other.
cookielawinfo-checkbox-performance11 monthsThis cookie is set by GDPR Cookie Consent plugin. The cookie is used to store the user consent for the cookies in the category "Performance".
viewed_cookie_policy11 monthsThe cookie is set by the GDPR Cookie Consent plugin and is used to store whether or not user has consented to the use of cookies. It does not store any personal data.

web counter

November 10, 2016

Exploring Enzymes

A catalyzing science project

By Science Buddies & Svenja Lohner

experiment about enzymes

Ready, set, react! Learn how enzymes power everything from our digestion to protecting our cells from damage--and watch them in action in this activity!

George Retseck

Key concepts Biology Biochemistry Enzymes Physiology Chemistry

Introduction Have you ever wondered how all the food that you eat gets digested? It is not only the acid in your stomach that breaks down your food—many little molecules in your body, called enzymes, help with that too. Enzymes are special types of proteins that speed up chemical reactions, such as the digestion of food in your stomach. In fact, there are thousands of different enzymes in your body that work around-the-clock to keep you healthy and active. In this science activity you will investigate one of these enzymes, called catalase, to find out how it helps to protect your body from damage.

Background Enzymes are essential for our survival. These proteins, made by our cells, help transform chemicals in our body, functioning as a catalyst. A catalyst gets reactions started and makes them happen faster, by increasing the rate of a reaction that otherwise might not happen at all, or would take too long to sustain life. However, a catalyst does not take part in the reaction itself—so how does this work? Each chemical reaction needs a minimum amount of energy to make it happen. This energy is called the activation energy. The lower the activation energy of a reaction, the faster it takes place. If the activation energy is too high, the reaction does not occur.

On supporting science journalism

If you're enjoying this article, consider supporting our award-winning journalism by subscribing . By purchasing a subscription you are helping to ensure the future of impactful stories about the discoveries and ideas shaping our world today.

Enzymes have the ability to lower the activation energy of a chemical reaction by interacting with its reactants (the chemicals doing the reacting). Each enzyme has an active site, which is where the reaction takes place. These sites are like special pockets that are able to bind a chemical molecule. The compounds or molecules the enzyme reacts with are called their substrates. The enzyme pocket has a special shape so that only one specific substrate is able to bind to it, just like only one key fits into a specific lock. Once the molecule is bound to the enzyme, the chemical reaction takes place. Then, the reaction products are released from the pocket, and the enzyme is ready to start all over again with another substrate molecule.

Catalase is a very common enzyme that is present in almost all organisms that are exposed to oxygen. The purpose of catalase in living cells is to protect them from oxidative damage, which can occur when cells or other molecules in the body come into contact with oxidative compounds. This damage is a natural result of reactions happening inside your cells. The reactions can include by-products such as hydrogen peroxide, which can be harmful to the body, just as how a by-product of a nice bonfire can be unwanted smoke that makes you cough or stings your eyes. To prevent such damage, the catalase enzyme helps getting rid of these compounds by breaking up hydrogen peroxide (H 2 O 2 ) into harmless water and oxygen. Do you want to see the catalyze enzyme in action? In this activity you will disarm hydrogen peroxide with the help of catalase from yeast.

Safety goggles or protective glasses

Five teaspoons of dish soap

One package of dry yeast

Hydrogen peroxide, 3 percent (at least 100 mL)

Three tablespoons

One teaspoon

Five 16-ounce disposable plastic cups

Measuring cup

Permanent marker

Paper towel

Workspace that can get wet (and won't be damaged by any spilled hydrogen peroxide or food-colored water)

Food coloring (optional)

Preparation

Take one cup and dissolve the dry yeast in about one-half cup of warm tap water. The water shouldn't be too hot but close to body temperature (37 Celsius). Let the dissolved yeast rest for at least five minutes.

Use the permanent marker to label the remaining four cups from one to four.

To all the labeled cups, add one teaspoon of dish soap.

To cup one no further additions are made at this point.

Before using the hydrogen peroxide, put on your safety goggles to protect your eyes. In case you spill hydrogen peroxide, clean it up with a wet paper towel. If you get it on your skin, make sure to rinse the affected area with plenty of water.

To cup two, add one tablespoon of 3 percent hydrogen peroxide solution. Use a fresh spoon for the hydrogen peroxide.

To cup three, add two tablespoons of  the hydrogen peroxide.

To cup four, add three tablespoons of the hydrogen peroxide.

Optionally, you can add a drop of food color to each of the labeled cups. (You can choose a different color for each one for easy identification)

Take cup number one and place it in front of you on the work area. With a fresh tablespoon, add one tablespoon of the dissolved yeast solution to the cup and swirl it slightly. What happens after you add the yeast? Do you see a reaction happening?

Place cup number two in front of you and again add one tablespoon of yeast solution to the cup. Once you add the enzyme, does the catalase react with the hydrogen peroxide? Can you see the reaction products being formed?

Add one tablespoon of yeast solution to cup number three. Do you see the same reaction taking place? Is the result different or the same compared to cup number two?

Finally, add one tablespoon of yeast solution to cup number four. Do you see more or less reaction products compared to your previous results? Can you explain the difference?

Place all four cups next to each other in front of you and observe your results. Did the enzymatic reaction take place in all of the cups or was there an exception? How do the results in each cup look different? Why do you think this is the case?

Now, take cup number one and add one additional tablespoon of 3 percent hydrogen peroxide to the cup. Swirl the cup slightly to mix the solution. What happens now? Looking at all your results, what do you think is the limiting factor for the catalase reaction in your cups?

Extra: Repeat this activity, but this time do not add dish soap to all of the reactions. What is different once you remove the dish soap? Do you still see foam formation?

Extra: So far you have observed the effect of substrate (H 2 O 2 ) concentration on the catalase reaction. What happens if you keep the substrate concentration constant but change the concentration of the enzyme? Try adding different amounts of yeast solution to three tablespoons of hydrogen peroxide, starting with one teaspoon. Do you observe any differences, or does the concentration of catalase not matter in your reaction?

Extra: What happens if the environmental conditions for the enzyme are changed? Repeat the catalase reaction but this time vary conditions such as the pH by adding vinegar (an acid) or baking soda (a base), or change the reaction temperature by heating the solution in the microwave. Can you identify which conditions are optimal for the catalase reaction? Are there any conditions that eliminate the catalase activity?

Extra: Can you find other sources of catalase enzyme that you could use in this activity? Research what other organisms, plants or cells contain catalase and try using these for your reaction. Do they work as well as yeast?

Observations and results You probably saw lots of bubbles and foam in this activity. What made the foam appear? When the enzyme catalase comes into contact with its substrate, hydrogen peroxide, it starts breaking it down into water and oxygen. Oxygen is a gas and therefore wants to escape the liquid. However, the dish soap that you added to all your solutions is able to trap the gas bubbles, which results in the formation of a stable foam. As long as there is enzyme and hydrogen peroxide present in the solution, the reaction continues and foam is produced. Once one of both compounds is depleted, the product formation stops. If you do not add dish soap to the reaction, you will see bubbles generated but no stable foam formation.

If there is no hydrogen peroxide present, the catalase cannot function, which is why in cup one you shouldn't have seen any bubble or foam production. Only when hydrogen peroxide is available, the catalase reaction can take place as you probably observed in the other cups. In fact, the catalase reaction is dependent on the substrate concentration. If you have an excess of enzyme but not enough substrate, the reaction will be limited by the substrate availability. Once you add more hydrogen peroxide to the solution, the reaction rate will increase as more substrate molecules can collide with the enzyme, forming more product. The result is an increasing amount of foam produced in your cup as you increase the amount of H 2 O 2 in your reaction. You should have seen more foam being produced once you added another tablespoon of hydrogen peroxide to cup one, which should have resulted in a similar amount of foam as in cup two. However, at some point you will reach a substrate concentration at which the enzyme gets saturated and becomes the limiting factor. In this case you have to add more enzyme to speed up the reaction again.

Many other factors affect the activity of enzymes as well. Most enzymes only function under optimal environmental conditions. If the pH or temperature deviates from these conditions too much, the enzyme reaction slows down significantly or does not work at all. You might have noticed that when doing the extra steps in the procedure.

Cleanup Pour all the solutions into the sink and clean all the spoons with warm water and dish soap. Wipe your work area with a wet paper towel and wash your hands with water and soap.

More to explore Biology for Kids; Enzymes , from Ducksters Enzymes: The Little Molecules That Bake Bread , from Scientific American Catalase , from PDB-101 Enzyme-Catalyzed Reactions—What Affects Their Rates? , from Science Buddies The Liver: Helping Enzymes Help You! , from Scientific American Science Activity for All Ages!, from Science Buddies

This activity brought to you in partnership with Science Buddies

experiment about enzymes

Optional Lab Activities

Lab objectives.

At the conclusion of the lab, the student should be able to:

  • define the following terms: metabolism, reactant, product, substrate, enzyme, denature
  • describe what the active site of an enzyme is (be sure to include information regarding the relationship of the active site to the substrate)
  • describe the specific action of the enzyme catalase, include the substrate and products of the reaction
  • list what organelle catalase can be found in every plant or animal cell
  • list the factors that can affect the rate of a chemical reaction and enzyme activity
  • explain why enzymes have an optimal pH and temperature to ensure greatest activity (greatest functioning) of the enzyme (be sure to consider how virtually all enzymes are proteins and the impact that temperature and pH may have on protein function)
  • explain why the same type of chemical reaction performed at different temperatures revealed different results/enzyme activity
  • explain why warm temperatures (but not boiling) typically promote enzyme activity but cold temperature typically
  • decreases enzyme activity
  • explain why increasing enzyme concentration promotes enzyme activity
  • explain why the optimal pH of a particular enzyme promotes its activity
  • if given the optimal conditions for a particular enzyme, indicate which experimental conditions using that particular enzyme would show the greatest and least enzyme activity

Introduction

Hydrogen peroxide is a toxic product of many chemical reactions that occur in living things. Although it is produced in small amounts, living things must detoxify this compound and break down hydrogen peroxide into water and oxygen, two non-harmful molecules. The organelle responsible for destroying hydrogen peroxide is the peroxisome using the enzyme catalase. Both plants and animals have peroxisomes with catalase. The catalase sample for today’s lab will be from a potato.

Enzymes speed the rate of chemical reactions. A catalyst is a chemical involved in, but not consumed in, a chemical reaction. Enzymes are proteins that catalyze biochemical reactions by lowering the activation energy necessary to break the chemical bonds in reactants and form new chemical bonds in the products. Catalysts bring reactants closer together in the appropriate orientation and weaken bonds, increasing the reaction rate. Without enzymes, chemical reactions would occur too slowly to sustain life.

The functionality of an enzyme is determined by the shape of the enzyme. The area in which bonds of the reactant(s) are broken is known as the active site. The reactants of enzyme catalyzed reactions are called substrates. The active site of an enzyme recognizes, confines, and orients the substrate in a particular direction.

Enzymes are substrate specific, meaning that they catalyze only specific reactions. For example, proteases (enzymes that break peptide bonds in proteins) will not work on starch (which is broken down by the enzyme amylase). Notice that both of these enzymes end in the suffix -ase. This suffix indicates that a molecule is an enzyme.

Environmental factors may affect the ability of enzymes to function. You will design a set of experiments to examine the effects of temperature, pH, and substrate concentration on the ability of enzymes to catalyze chemical reactions. In particular, you will be examining the effects of these environmental factors on the ability of catalase to convert H 2 O 2 into H 2 O and O 2 .

The Scientific Method

As scientists, biologists apply the scientific method. Science is not simply a list of facts, but is an approach to understanding the world around us. It is use of the scientific method that differentiates science from other fields of study that attempt to improve our understanding of the world.

The scientific method is a systematic approach to problem solving. Although some argue that there is not one single scientific method, but a variety of methods; each of these approaches, whether explicit or not, tend to incorporate a few fundamental steps: observing, questioning, hypothesizing, predicting, testing, and interpreting results of the test. Sometimes the distinction between these steps is not always clear. This is particularly the case with hypotheses and predictions. But for our purposes, we will differentiate each of these steps in our applications of the scientific method.

You are already familiar with the steps of the scientific method from previous lab experiences. You will need to use your scientific method knowledge in today’s lab in creating hypotheses for each experiment, devising a protocol to test your hypothesis, and analyzing the results. Within the experimentation process it will be important to identify the independent variable, the dependent variable, and standardized variables for each experiment.

Part 1: Observe the Effects of Catalase

  • Obtain two test tubes and label one as A and one as B.
  • Use your ruler to measure and mark on each test tube 1 cm from the bottom.
  • Fill each of two test tubes with catalase (from the potato) to the 1 cm mark
  • Add 10 drops of hydrogen peroxide to the tube marked A.
  • Add 10 drops of distilled water to the tube marked B.
  • Bubbling height tube A
  • Bubbling height tube B
  • What happened when H 2 O 2 was added to the potato in test tube A?
  • What caused this to happen?
  • What happened in test tube B?
  • What was the purpose of the water in tube B?

Part 2: Effects of pH, Temperature, and Substrate Concentration

Observations.

From the introduction and your reading, you have some background knowledge on enzyme structure and function. You also just observed the effects of catalase on the reaction in which hydrogen peroxide breaks down into water and oxygen.

From the objectives of this lab, our questions are as follows:

  • How does temperature affect the ability of enzymes to catalyze chemical reactions?
  • How does pH affect the ability of enzymes to catalyze chemical reactions?
  • What is the effect of substrate concentration on the rate of enzyme catalyzed reactions?

Based on the questions above, come up with some possible hypotheses. These should be general, not specific, statements that are possible answers to your questions.

  • Temperature hypothesis
  • pH hypothesis
  • Substrate concentration hypothesis

Test Your Hypotheses

Based on your hypotheses, design a set of experiments to test your hypotheses. Use your original experiment to shape your ideas. You have the following materials available:

  • Catalase (from potato)
  • Hydrogen peroxide
  • Distilled water
  • Hot plate (for boiling water)
  • Acidic pH solution
  • Basic pH solution
  • Thermometer
  • Ruler and wax pencil

Write your procedure to test each hypothesis. You should have three procedures, one for each hypothesis. Make sure your instructor checks your procedures before you continue.

  • Procedure 1: Temperature
  • Procedure 2: pH
  • Procedure 3: Concentration

Record your results—you may want to draw tables. Also record any observations you make. Interpret your results to draw conclusions.

  • Do your results match your hypothesis for each experiment?
  • Do the results reject or fail to reject your hypothesis and why?
  • What might explain your results? If your results are different from your hypothesis, why might they differ? If the results matched your predictions, hypothesize some mechanisms behind what you have observed.

Communicating Your Findings

Scientists generally communicate their research findings in written reports. Save the things that you have done above. You will be use them to write a lab report a little later in the course.

Sections of a Lab Report

  • Title Page:  The title describes the focus of the research. The title page should also include the student’s name, the lab instructor’s name, and the lab section.
  • Introduction:  The introduction provides the reader with background information about the problem and provides the rationale for conducting the research. The introduction should incorporate and cite outside sources. You should avoid using websites and encyclopedias for this background information. The introduction should start with more broad and general statements that frame the research and become more specific, clearly stating your hypotheses near the end.
  • Methods:  The methods section describes how the study was designed to test your hypotheses. This section should provide enough detail for someone to repeat your study. This section explains what you did. It should not be a bullet list of steps and materials used; nor should it read like a recipe that the reader is to follow. Typically this section is written in first person past tense in paragraph form since you conducted the experiment.
  • Results:  This section provides a written description of the data in paragraph form. What was the most reaction? The least reaction? This section should also include numbered graphs or tables with descriptive titles. The objective is to present the data, not interpret the data. Do not discuss why something occurred, just state what occurred.
  • Discussion:  In this section you interpret and critically evaluate your results. Generally, this section begins by reviewing your hypotheses and whether your data support your hypotheses. In describing conclusions that can be drawn from your research, it is important to include outside studies that help clarify your results. You should cite outside resources. What is most important about the research? What is the take-home message? The discussion section also includes ideas for further research and talks about potential sources of error. What could you improve if you conducted this experiment a second time?
  • Biology 101 Labs. Authored by : Lynette Hauser. Provided by : Tidewater Community College. Located at : http://www.tcc.edu/ . License : CC BY: Attribution
  • BIOL 160 - General Biology with Lab. Authored by : Scott Rollins. Provided by : Open Course Library. Located at : http://opencourselibrary.org/biol-160-general-biology-with-lab/ . License : CC BY: Attribution

Your browser is not supported

Sorry but it looks as if your browser is out of date. To get the best experience using our site we recommend that you upgrade or switch browsers.

Find a solution

  • Skip to main content
  • Skip to navigation

experiment about enzymes

  • Back to parent navigation item
  • Primary teacher
  • Secondary/FE teacher
  • Early career or student teacher
  • Higher education
  • Curriculum support
  • Literacy in science teaching
  • Periodic table
  • Interactive periodic table
  • Climate change and sustainability
  • Resources shop
  • Collections
  • Remote teaching support
  • Starters for ten
  • Screen experiments
  • Assessment for learning
  • Microscale chemistry
  • Faces of chemistry
  • Classic chemistry experiments
  • Nuffield practical collection
  • Anecdotes for chemistry teachers
  • On this day in chemistry
  • Global experiments
  • PhET interactive simulations
  • Chemistry vignettes
  • Context and problem based learning
  • Journal of the month
  • Chemistry and art
  • Art analysis
  • Pigments and colours
  • Ancient art: today's technology
  • Psychology and art theory
  • Art and archaeology
  • Artists as chemists
  • The physics of restoration and conservation
  • Ancient Egyptian art
  • Ancient Greek art
  • Ancient Roman art
  • Classic chemistry demonstrations
  • In search of solutions
  • In search of more solutions
  • Creative problem-solving in chemistry
  • Solar spark
  • Chemistry for non-specialists
  • Health and safety in higher education
  • Analytical chemistry introductions
  • Exhibition chemistry
  • Introductory maths for higher education
  • Commercial skills for chemists
  • Kitchen chemistry
  • Journals how to guides
  • Chemistry in health
  • Chemistry in sport
  • Chemistry in your cupboard
  • Chocolate chemistry
  • Adnoddau addysgu cemeg Cymraeg
  • The chemistry of fireworks
  • Festive chemistry
  • Education in Chemistry
  • Teach Chemistry
  • On-demand online
  • Live online
  • Selected PD articles
  • PD for primary teachers
  • PD for secondary teachers
  • What we offer
  • Chartered Science Teacher (CSciTeach)
  • Teacher mentoring
  • UK Chemistry Olympiad
  • Who can enter?
  • How does it work?
  • Resources and past papers
  • Top of the Bench
  • Schools' Analyst
  • Regional support
  • Education coordinators
  • RSC Yusuf Hamied Inspirational Science Programme
  • RSC Education News
  • Supporting teacher training
  • Interest groups

A primary school child raises their hand in a classroom

  • More navigation items

Testing for catalase enzymes

In association with Nuffield Foundation

  • No comments

Try this class experiment to detect the presence of enzymes as they catalyse the decomposition of hydrogen peroxide

Enzymes are biological catalysts which increase the speed of a chemical reaction. They are large protein molecules and are very specific to certain reactions. Hydrogen peroxide decomposes slowly in light to produce oxygen and water. The enzyme catalase can speed up (catalyse) this reaction.

In this practical, students investigate the presence of enzymes in liver, potato and celery by detecting the oxygen gas produced when hydrogen peroxide decomposes. The experiment should take no more than 20–30 minutes.

  • Eye protection
  • Conical flasks, 100 cm 3 , x3
  • Measuring cylinder, 25 cm 3
  • Bunsen burner
  • Wooden splint
  • A bucket or bin for disposal of waste materials
  • Hydrogen peroxide solution, ‘5 volume’

Health, safety and technical notes

  • Read our standard health and safety guidance.
  • Wear eye protection throughout. Students must be instructed NOT to taste or eat any of the foods used in the experiment.
  • Hydrogen peroxide solution, H 2 O 2 (aq) – see CLEAPSS Hazcard HC050  and CLEAPSS Recipe Book RB045. Hydrogen peroxide solution of ‘5 volume’ concentration is low hazard, but it will probably need to be prepared by dilution of a more concentrated solution which may be hazardous.
  • Only small samples of liver, potato and celery are required. These should be prepared for the lesson ready to be used by students. A disposal bin or bucket for used samples should be provided to avoid these being put down the sink.
  • Measure 25 cm 3  of hydrogen peroxide solution into each of three conical flasks.
  • At the same time, add a small piece of liver to the first flask, a small piece of potato to the second flask, and a small piece of celery to the third flask.
  • Hold a glowing splint in the neck of each flask.
  • Note the time taken before each glowing splint is relit by the evolved oxygen.
  • Dispose of all mixtures into the bucket or bin provided.

Teaching notes

Some vegetarian students may wish to opt out of handling liver samples, and this should be respected.

Before or after the experiment, the term enzyme will need to be introduced. The term may have been met previously in biological topics, but the notion that they act as catalysts and increase the rate of reactions may be new. Similarly their nature as large protein molecules whose catalytic activity can be very specific to certain chemical reactions may be unfamiliar. The name catalase for the enzyme present in all these foodstuffs can be introduced.

To show the similarity between enzymes and chemical catalysts, the teacher may wish to demonstrate (or ask the class to perform as part of the class experiment) the catalytic decomposition of hydrogen peroxide solution by manganese(IV) oxide (HARMFUL – see CLEAPSS Hazcard HC060).

If students have not performed the glowing splint test for oxygen for some time, they may need reminding of how to do so by a quick demonstration by the teacher.

More resources

Add context and inspire your learners with our short career videos showing how chemistry is making a difference .

Additional information

This is a resource from the  Practical Chemistry project , developed by the Nuffield Foundation and the Royal Society of Chemistry.

Practical Chemistry activities accompany  Practical Physics  and  Practical Biology .

© Nuffield Foundation and the Royal Society of Chemistry

  • 11-14 years
  • 14-16 years
  • 16-18 years
  • Practical experiments
  • Biological chemistry

Specification

  • Enzymes act as catalysts in biological systems.
  • Factors which affect the rates of chemical reactions include: the concentrations of reactants in solution, the pressure of reacting gases, the surface area of solid reactants, the temperature and the presence of catalysts.
  • Describe the characteristics of catalysts and their effect on rates of reaction.
  • Recall that enzymes act as catalysts in biological systems.
  • 7.6 Describe a catalyst as a substance that speeds up the rate of a reaction without altering the products of the reaction, being itself unchanged chemically and in mass at the end of the reaction
  • 7.8 Recall that enzymes are biological catalysts and that enzymes are used in the production of alcoholic drinks
  • C6.2.4 describe the characteristics of catalysts and their effect on rates of reaction
  • C6.2.5 identify catalysts in reactions
  • C6.2.14 describe the use of enzymes as catalysts in biological systems and some industrial processes
  • C5.2f describe the characteristics of catalysts and their effect on rates of reaction
  • C5.2i recall that enzymes act as catalysts in biological systems
  • C6.2.13 describe the use of enzymes as catalysts in biological systems and some industrial processes
  • C5.1f describe the characteristics of catalysts and their effect on rates of reaction
  • C5.1i recall that enzymes act as catalysts in biological systems
  • B2.24 The action of a catalyst, in terms of providing an alternative pathway with a lower activation energy.
  • 2.3.5 demonstrate knowledge and understanding that a catalyst is a substance which increases the rate of a reaction without being used up and recall that transition metals and their compounds are often used as catalysts;
  • 7. Investigate the effect of a number of variables on the rate of chemical reactions including the production of common gases and biochemical reactions.

Related articles

Students in a school lab using a burette filler to measure liquid for a titration

How to teach titration post-16

2024-07-08T05:32:00Z By Jo Haywood

Tips for teaching practical titration techniques and the underlying theory

Stacked spheres showing the molecular structure of a diamond

New method grows larger diamonds

2024-05-31T08:24:00Z By Nina Notman

Use this real-world context when teaching about giant covalent structures

A dog playing in a field of narcissus daffodil flowers

Why are some plants poisonous to you and your pets?

2024-05-22T08:16:00Z By Kit Chapman

Dig up the toxic secrets of nature’s blooms

No comments yet

Only registered users can comment on this article., more experiments.

Image showing a one page from the technician notes, teacher notes, student sheet and integrated instructions that make up this resource, plus two bags of chocolate coins

‘Gold’ coins on a microscale | 14–16 years

By Dorothy Warren and Sandrine Bouchelkia

Practical experiment where learners produce ‘gold’ coins by electroplating a copper coin with zinc, includes follow-up worksheet

potion labels

Practical potions microscale | 11–14 years

By Kirsty Patterson

Observe chemical changes in this microscale experiment with a spooky twist.

An image showing the pages available in the downloads with a water bottle in the shape of a 6 in the foreground.

Antibacterial properties of the halogens | 14–18 years

By Kristy Turner

Use this practical to investigate how solutions of the halogens inhibit the growth of bacteria and which is most effective

  • Contributors
  • Email alerts

Site powered by Webvision Cloud

menu

Search My Blog

Everything About Enzymes!! (and a free lab!)

experiment about enzymes

  • Enzymes are biological catalysts that speed up the chemical reactions of the cell.
  • Enzymes are proteins.
  • Enzymatic reactions occur faster and at lower temperatures because enzymes lower the activation energy for that chemical reaction.
  • Enzymes are never consumed or used up during the reaction. They can do their job over and over again.

experiment about enzymes

  • Enzymes are highly specific for just one substrate.  The enzyme has an active site with a unique 3-D shape into which this substrate must fit.  
  • Enzymes catalyze both the forward and the reverse of the same reaction.
  • Enzymes can be denatured by temperatures and pH levels outside the optimal range for that particular enzyme.

experiment about enzymes

I teach 2nd grade during the year, but do high school science camps in the summer. Can't wait to check these out and see if they work for us when we're studying enzymes! Thanks! Jenny Luckeyfrog's Lilypad

Investigation: Enzymes

download pdf

Enzyme Lab Teacher's Guide

Measure the effects of changes in temperature, pH, and enzyme concentration on reaction rates of an enzyme

Explain how environmental factors affect the rate of enzyme-catalyzed reactions.

INTRODUCTION: What would happen to your cells if they made a poisonous chemical? You might think that they would die. In fact, your cells are always making poisonous chemicals. They do not die because your cells use enzymes to break down these poisonous chemicals into harmless substances.

Enzymes are proteins that speed up the rate of reactions that would otherwise happen more slowly. The enzyme is not altered by the reaction. You have hundreds of different enzymes in each of your cells.

Each of these enzymes is responsible for one particular reaction that occurs in the cell. In this lab, you will study an enzyme that is found in the cells of many living tissues. The name of the enzyme is catalase; it speeds up a reaction which breaks down hydrogen peroxide, a toxic chemical, into 2 harmless substances--water and oxygen. Light can also break down H 2 O 2 which is why the chemical is sold in dark containers.

The reaction is: 2H 2 O 2 → 2H 2 O + O 2

This reaction is important to cells because hydrogen peroxide is produced as a byproduct of many normal cellular reactions. If the cells did not break down the hydrogen peroxide, they would be poisoned and die. In this lab, you will study the catalase found in liver cells. You will be using chicken or beef liver. It might seem strange to use dead cells to study the function of enzymes. This is possible because when a cell dies, the enzymes remain intact and active for several weeks, as long as the tissue is kept refrigerated.

MATERIALS: 6 Test tubes / Test tube holders 3% Hydrogen peroxide Ice / Hot water Straight-edged razor blade  Scissors and Forceps Measuring Pipettes Stirring rod 

Fresh liver, Apple, and Potato, Yeast Vinegar / Baking Soda HCL and NaOH pH paper (optional)

PART A - Observe Normal Catalase Reaction

1. Place about 2 ml of the 3% hydrogen peroxide solution into a clean test tube.

2. Using forceps and scissors cut a small piece of liver and add it to the test tube. Push it into the hydrogen peroxide with a stirring rod. Observe the bubbles.         What gas is being released? (consider the equation above) _____

Throughout this investigation you will estimate the rate of the reaction (how rapidly the solution bubbles) on a scale of 0-5 (0=no reaction, 1=slow, ..... 5= very fast). Assume that the reaction in step 2 proceeded at a rate of "4"

Recall that a reaction that absorbs heat is endothermic; a reaction that gives off heat is exothermic. Feel the temperature of the test tube with your hand.

Has it gotten warmer or colder? Is the reaction endothermic or exothermic?

3. Pour off the liquid into a second test tube.            Assuming the reaction is complete, what is this liquid composed of?

What do you think would happen if you added more liver to this liquid?

Test this and record the reaction rate.  Reaction Rate:

4.  Add another 2ml of hydrogen peroxide to the liver remaining in the first test tube.   What is the reaction rate?

Synthesis --  Answer the question:  Is catalase reusable?   

REASONING .  

Part B - What Tissues Contain Catalase?

You will now test for the presence of catalase in tissues other than liver. Place 2 ml of hydrogen peroxide in each of 3 clean test tubes and then add each of the three test substances to the tubes.  As you add each test substance, record the reaction rate (0-5) for each tube.

Substance

Apple

Potato

Yeast

Rate of Reaction (0-5)

 

 

 

water bath

Synthesis -- Do all living tissues contain catalase?   Claim:                                         

Evidence: Reasoning:

test tubes

PART C - What is the Effect of Temperature on Catalase Activity?

1. Put a piece of liver into the bottom of a clean test tube and cover it with a small amount of water. Place this test tube in a boiling water bath for 5 minutes.

2. Remove the test tube from the hot water bath, allow it to air cool, then pour out the water. Add 2 ml of hydrogen peroxide. CAUTION: Use a test-tube holder for hot test tubes. 

 What is the reaction rate for the boiled liver and peroxide? __________

3. Put equal quantities of liver into 2 clean test tubes and 1 ml H 2 O 2 into 2 other test tubes. Put one test tube of liver and one of H 2 O 2 into an ice bath. Place the other set in a warm water bath (not boiling).

After 3 minutes, pour each tube of H 2 O 2 into the corresponding tube of liver and observe the reaction

What is the reaction rate for the cold liver/peroxide? _____  What is the reaction rate for the warm liver/peroxide? ____

Synthesis -- How does temperature affect the catalase enzyme?

Claim:                             

PART D - What is the Effect of pH on Catalase Activity

1. Add 2 ml hydrogen peroxide to 4 clean test tubes, then add:

Tube 1 – add 3 drops of acetic acid (vinegar)  pH =_______  Tube 2 – add 3 drops of sodium bicarbonate (base) pH =______ Tube 3 – add 3 drops of water (neutral) pH =_____ Tube 4 -- add 3 drops of 1M NaOH   pH =  _____

Now add liver to each of the test tubes (try to do it all at about the same time, so you can easily compare)

Rate of Reaction for:

Strong Acid  (HCL)  ____     Acid _____     Neutral ______   Base_____     Strong base (NAOH) _____

1.  How does  pH affect the reaction rate of catalase? Propose a way to  refine  your experiment to find the  exact , or OPTIMAL pH and temperature of catalase.

2.  The following graph shows reaction rates of various enzymes in the body.  Pepsin is found in the stomach, amylase in the saliva, and phosphatase in the liver. 

graph

Synthesis:  How does pH affect the activity of enzymes?

Claim:      

                                        

lactaid

Part E - Design an Experiment

Lactaid is a product designed to help people who cannot digest milk sugar (lactose) because they are missing the enzyme lactase. Many people are lactose-intolerant, a condition that is mainly genetic. Lactase breaks down lactose into two subunits: glucose and galactose.

To test for the presence of monosaccharides and reducing disaccharide sugars in food, the food sample is dissolved in water, and a small amount of Benedict's reagent is added. The solution should progress in the colors of blue (with no glucose present), green, yellow, orange, red, and then brick red when there is a large amount of glucose present. (Google benedict's test to see the way this looks.)

Design an experiment where you would determine how quicly lactaid works to break down milk sugar at different temperatures.. Be specific in your description, use drawings if necessary.

Other Resources on Enzymes

Analyzing Graphics - Enzymes - shows substrates and enzyme interactions and explores competitive inhibition

Observe Catalase Activity in Yeast - create sodium alginate spheres to observe how catalase breaks down hydrogen peroxide

Enzyme Activity Using Toothpickase - simulate the activity of an enzyme by breaking toothpicks.

Laboratory Manual For SCI103 Biology I at Roxbury Community College

Enzymes are macromolecular biological catalysts. The molecules upon which enzymes may act are called substrates and the enzyme converts the substrates into different molecules known as products. Almost all metabolic processes in the cell need enzyme catalysis in order to occur at rates fast enough to sustain life. Metabolic pathways depend upon enzymes to catalyze individual steps. Enzymes are known to catalyze more than 5,000 biochemical reaction types. Most enzymes are proteins, although a few are catalytic RNA molecules. The latter are called ribozymes. Enzymes’ specificity comes from their unique three-dimensional structures.

Like all catalysts, enzymes increase the reaction rate by lowering its activation energy. Some enzymes can make their conversion of substrate to product occur many millions of times faster. An extreme example is orotidine 5’-phosphate decarboxylase, which allows a reaction that would otherwise take millions of years to occur in milliseconds. Chemically, enzymes are like any catalyst and are not consumed in chemical reactions, nor do they alter the equilibrium of a reaction. Enzymes differ from most other catalysts by being much more specific. Enzyme activity can be affected by other molecules: inhibitors are molecules that decrease enzyme activity, and activators are molecules that increase activity. Many therapeutic drugs and poisons are enzyme inhibitors. An enzyme’s activity decreases markedly outside its optimal temperature and pH.

Some enzymes are used commercially, for example, in the synthesis of antibiotics. Some household products use enzymes to speed up chemical reactions: enzymes in biological washing powders break down protein, starch or fat stains on clothes, and enzymes in meat tenderizer break down proteins into smaller molecules, making the meat easier to chew.

In this laboratory, we will study the effect of temperature, concentration and pH and on the activity of the enzyme catalase. Catalase speeds up the following reaction:

2 H 2 O 2 -> 2 H 2 O + O 2

Hydrogen peroxide is toxic. Cells therefore use catalase to protect themselves. In these experiments, we will use catalase enzyme from potato.

The first experiment will establish that our catalase works (positive control) and that our reagents are not contaminated (negative control).

Hydrogen peroxide will not spontaneously degrade at room temperature in the absence of enzyme. When catalase is added to hydrogen peroxide, the reaction will take place and the oxygen produced will lead to the formation of bubbles in the solution. The height of the bubbles above the solution will be our measure of enzyme activity (Figure 8.2 ).

8.1 Positive and negative controls (Experiment 1)

8.1.1 experimental procedures.

  • Obtain and label three 15 ml conical plastic reaction tubes.
  • Add 1 ml of potato juice (catalase) to tube 1 (use a plastic transfer pipette).
  • Add 4 ml of hydrogen peroxide to tube 1. Swirl well to mix and wait at least 20 seconds for bubbling to develop.
  • Use a ruler (Figure 8.1 ) and measure the height of the bubble column above the liquid (in millimeters; use the centimeter scale of the ruler) and record the result in Table 8.1 .
  • Add 1 ml of water.
  • Add 4 ml of hydrogen peroxide. Swirl well to mix and wait at least 20 seconds.
  • Measure the height of the bubble column (in millimeters) and record the result in Table 8.1 .
  • Add 1 ml of potato juice (catalase).
  • Add 4 ml of sucrose solution. Swirl well to mix; wait 20 seconds.
  • Measure the height of the bubble column and record the result in Table 8.1 .
Table 8.1: Positive and negative controls.
Tube # Height of bubbles (mm)
1
2
3

A ruler with metric (cm) and imperial (inch) scales.

Figure 8.1: A ruler with metric (cm) and imperial (inch) scales.

Results from experiment 1. Compare with your results!

Figure 8.2: Results from experiment 1. Compare with your results!

8.2 Effect of temperature on enzyme activity (Experiment 2)

8.2.1 experimental procedures.

Before you begin with the actual experiment, write down in your own words the hypothesis for this experiment:

  • Obtain and label three tubes.
  • Add 1 ml of potato juice (catalase) to each tube.
  • Place tube 1 in the refrigerator, tube 2 in a 37 °C (Celsius) heat block, and tube 3 in a 97 °C heat block for 15 minutes.
  • Remove the tubes with the potato juice (catalase) from the refrigerator and heat blocks and immediately add 4 ml hydrogen peroxide to each tube.
  • Swirl well to mix and wait 20 seconds.
  • Measure the height of the bubble column (in millimeters) in each tube and record your observations in Table 8.2 .

Do the data support or contradict your hypothesis?

Table 8.2: Effect of temperature on enzyme activity.
Tube # Height of bubbles (mm)
1
2
3

Results from experiment 2. Compare with your results!

Figure 8.3: Results from experiment 2. Compare with your results!

8.3 Effect of concentration on enzyme activity (Experiment 3)

8.3.1 experimental procedures.

  • Measure the height of the bubble column (in millimeters) and record your observations in Table 8.3 .
  • Add 3 ml of potato juice (catalase).
Table 8.3: Effect of concentration on enzyme activity.
Tube # Height of bubbles (mm)
1
2
3

Results from experiment 3. Compare with your results!

Figure 8.4: Results from experiment 3. Compare with your results!

8.4 Effect of pH on enzyme activity (Experiment 4)

8.4.1 experimental procedures.

  • Obtain 6 tubes and label each tube with a number from 1 to 6.
  • Place the tubes from left (tube #1) to right (tube #6) in the first row of a test tube rack.
  • Add to 1 ml of potato juice (catalase) to each tube.
  • Add 2 ml of water to tube 1.
  • Add 2 ml of pH buffer 3 to tube 2.
  • Add 2 ml of pH buffer 5 to tube 3.
  • Add 2 ml of pH buffer 7 to tube 4.
  • Add 2 ml of pH buffer 9 to tube 5.
  • Add 2 ml of pH buffer 12 to tube 6.
  • Add 4 ml of hydrogen peroxide to each of the six tubes.
  • Swirl each tube well to mix and wait at least 20 seconds.
  • Measure the height of the bubble column (in millimeters) in each tube and record your observations in Table 8.4 .
Table 8.4: Effect of pH on enzyme activity.
Tube # Height of bubbles (mm)
1
2
3
4
5
6

Results from experiment 4. Compare with your results!

Figure 8.5: Results from experiment 4. Compare with your results!

Figure 8.6: Catalase activity is dependent on pH. The data shown in this figure were obtained by three groups of students during a previous laboratory session. The triangles represent the data from the experimental results shown in Figure 8.5 .

8.5 Cleaning up

  • Empty the contents of the plastic tubes into the labeled waste container (brown bottle) in the chemical fume hood.
  • Discard the empty tubes and other waste in the regular waste basket.
  • Rinse the glass rod and glassware with water and detergent.
  • Return the glass ware to the trays on your bench where you originally found them.

8.6 Review Questions

  • What is a catalyst?
  • What are enzymes?
  • What is the name of the enzyme that we studied in this laboratrory session?
  • What is an enzyme substrate?
  • What is the substrate of the enzyme that we used in this laborator seesion?
  • What are the products of the reaction that was catalized by the enzyme that we studied in this laboratory seesion?
  • What is the active site of an enzyme?
  • What is the purpose of the negative and positive controls?
  • State the hypothesis that was tested in experiment 2?
  • State the hypothesis that was tested in experiment 3?
  • State the hypothesis that was tested in experiment 4?
  • The enzyme from potato appeared to work better at 4 °C than at 37 °C. Would you expect the same if we had used the equivalent human enzyme? Justify your answer.
  • Why did heating the enzyme at high temperature (> 65 °C) result in loss of activity?

U.S. flag

An official website of the United States government

The .gov means it’s official. Federal government websites often end in .gov or .mil. Before sharing sensitive information, make sure you’re on a federal government site.

The site is secure. The https:// ensures that you are connecting to the official website and that any information you provide is encrypted and transmitted securely.

  • Publications
  • Account settings

Preview improvements coming to the PMC website in October 2024. Learn More or Try it out now .

  • Advanced Search
  • Journal List
  • Portland Press Open Access

Logo of portlandopen

Enzymes: principles and biotechnological applications

Enzymes are biological catalysts (also known as biocatalysts) that speed up biochemical reactions in living organisms, and which can be extracted from cells and then used to catalyse a wide range of commercially important processes. This chapter covers the basic principles of enzymology, such as classification, structure, kinetics and inhibition, and also provides an overview of industrial applications. In addition, techniques for the purification of enzymes are discussed.

The nature and classification of enzymes

Enzymes are biological catalysts (also known as biocatalysts) that speed up biochemical reactions in living organisms. They can also be extracted from cells and then used to catalyse a wide range of commercially important processes. For example, they have important roles in the production of sweetening agents and the modification of antibiotics, they are used in washing powders and various cleaning products, and they play a key role in analytical devices and assays that have clinical, forensic and environmental applications. The word ‘enzyme’ was first used by the German physiologist Wilhelm Kühne in 1878, when he was describing the ability of yeast to produce alcohol from sugars, and it is derived from the Greek words en (meaning ‘within’) and zume (meaning ‘yeast’).

In the late nineteenth century and early twentieth century, significant advances were made in the extraction, characterization and commercial exploitation of many enzymes, but it was not until the 1920s that enzymes were crystallized, revealing that catalytic activity is associated with protein molecules. For the next 60 years or so it was believed that all enzymes were proteins, but in the 1980s it was found that some ribonucleic acid (RNA) molecules are also able to exert catalytic effects. These RNAs, which are called ribozymes, play an important role in gene expression. In the same decade, biochemists also developed the technology to generate antibodies that possess catalytic properties. These so-called ‘abzymes’ have significant potential both as novel industrial catalysts and in therapeutics. Notwithstanding these notable exceptions, much of classical enzymology, and the remainder of this essay, is focused on the proteins that possess catalytic activity.

As catalysts, enzymes are only required in very low concentrations, and they speed up reactions without themselves being consumed during the reaction. We usually describe enzymes as being capable of catalysing the conversion of substrate molecules into product molecules as follows:

Enzymes are potent catalysts

The enormous catalytic activity of enzymes can perhaps best be expressed by a constant, k cat , that is variously referred to as the turnover rate, turnover frequency or turnover number. This constant represents the number of substrate molecules that can be converted to product by a single enzyme molecule per unit time (usually per minute or per second). Examples of turnover rate values are listed in Table 1 . For example, a single molecule of carbonic anhydrase can catalyse the conversion of over half a million molecules of its substrates, carbon dioxide (CO 2 ) and water (H 2 O), into the product, bicarbonate (HCO 3 − ), every second—a truly remarkable achievement.

EnzymeTurnover rate (mole product s mole enzyme )
Carbonic anhydrase600 000
Catalase93 000
β–galactosidase200
Chymotrypsin100
Tyrosinase1

Enzymes are specific catalysts

As well as being highly potent catalysts, enzymes also possess remarkable specificity in that they generally catalyse the conversion of only one type (or at most a range of similar types) of substrate molecule into product molecules.

Some enzymes demonstrate group specificity. For example, alkaline phosphatase (an enzyme that is commonly encountered in first-year laboratory sessions on enzyme kinetics) can remove a phosphate group from a variety of substrates.

Other enzymes demonstrate much higher specificity, which is described as absolute specificity. For example, glucose oxidase shows almost total specificity for its substrate, β-D-glucose, and virtually no activity with any other monosaccharides. As we shall see later, this specificity is of paramount importance in many analytical assays and devices (biosensors) that measure a specific substrate (e.g. glucose) in a complex mixture (e.g. a blood or urine sample).

Enzyme names and classification

Enzymes typically have common names (often called ‘trivial names’) which refer to the reaction that they catalyse, with the suffix -ase (e.g. oxidase, dehydrogenase, carboxylase), although individual proteolytic enzymes generally have the suffix - in (e.g. trypsin, chymotrypsin, papain). Often the trivial name also indicates the substrate on which the enzyme acts (e.g. glucose oxidase, alcohol dehydrogenase, pyruvate decarboxylase). However, some trivial names (e.g. invertase, diastase, catalase) provide little information about the substrate, the product or the reaction involved.

Due to the growing complexity of and inconsistency in the naming of enzymes, the International Union of Biochemistry set up the Enzyme Commission to address this issue. The first Enzyme Commission Report was published in 1961, and provided a systematic approach to the naming of enzymes. The sixth edition, published in 1992, contained details of nearly 3 200 different enzymes, and supplements published annually have now extended this number to over 5 000.

Within this system, all enzymes are described by a four-part Enzyme Commission (EC) number. For example, the enzyme with the trivial name lactate dehydrogenase has the EC number 1.1.1.27, and is more correctly called l –lactate: NAD + oxidoreductase.

The first part of the EC number refers to the reaction that the enzyme catalyses ( Table 2 ). The remaining digits have different meanings according to the nature of the reaction identified by the first digit. For example, within the oxidoreductase category, the second digit denotes the hydrogen donor ( Table 3 ) and the third digit denotes the hydrogen acceptor ( Table 4 ).

First EC digitEnzyme classReaction type
1.OxidoreductasesOxidation/reduction
2.TransferasesAtom/group transfer (excluding other classes)
3.HydrolasesHydrolysis
4.LyasesGroup removal (excluding 3.)
5.IsomerasesIsomerization
6.LigasesJoining of molecules linked to the breakage of a pyrophosphate bond
Oxidoreductases: second EC digitHydrogen or electron donor
1.Alcohol (CHOH)
2.Aldehyde or ketone (C═O)
3.─CH─CH─
4.Primary amine (CHNH or CHNH )
5.Secondary amine (CHNH)
6.NADH or NADPH (when another redox catalyst is the acceptor)
Oxidoreductases: third EC digitHydrogen or electron acceptor
1.NAD or NADP
2.Fe (e.g. cytochromes)
3.O
4.Other

Thus lactate dehydrogenase with the EC number 1.1.1.27 is an oxidoreductase (indicated by the first digit) with the alcohol group of the lactate molecule as the hydrogen donor (second digit) and NAD + as the hydrogen acceptor (third digit), and is the 27th enzyme to be categorized within this group (fourth digit).

An external file that holds a picture, illustration, etc.
Object name is bse0590001in01.jpg

Fortunately, it is now very easy to find this information for any individual enzyme using the Enzyme Nomenclature Database (available at http://enzyme.expasy.org ).

Enzyme structure and substrate binding

Amino acid-based enzymes are globular proteins that range in size from less than 100 to more than 2 000 amino acid residues. These amino acids can be arranged as one or more polypeptide chains that are folded and bent to form a specific three-dimensional structure, incorporating a small area known as the active site ( Figure 1 ), where the substrate actually binds. The active site may well involve only a small number (less than 10) of the constituent amino acids.

An external file that holds a picture, illustration, etc.
Object name is bse0590001i001.jpg

It is the shape and charge properties of the active site that enable it to bind to a single type of substrate molecule, so that the enzyme is able to demonstrate considerable specificity in its catalytic activity.

The hypothesis that enzyme specificity results from the complementary nature of the substrate and its active site was first proposed by the German chemist Emil Fischer in 1894, and became known as Fischer's ‘lock and key hypothesis’, whereby only a key of the correct size and shape (the substrate) fits into the keyhole (the active site) of the lock (the enzyme). It is astounding that this theory was proposed at a time when it was not even established that enzymes were proteins. As more was learned about enzyme structure through techniques such as X-ray crystallography, it became clear that enzymes are not rigid structures, but are in fact quite flexible in shape. In the light of this finding, in 1958 Daniel Koshland extended Fischer's ideas and presented the ‘induced-fit model’ of substrate and enzyme binding, in which the enzyme molecule changes its shape slightly to accommodate the binding of the substrate. The analogy that is commonly used is the ‘hand-in-glove model’, where the hand and glove are broadly complementary in shape, but the glove is moulded around the hand as it is inserted in order to provide a perfect match.

Since it is the active site alone that binds to the substrate, it is logical to ask what is the role of the rest of the protein molecule. The simple answer is that it acts to stabilize the active site and provide an appropriate environment for interaction of the site with the substrate molecule. Therefore the active site cannot be separated out from the rest of the protein without loss of catalytic activity, although laboratory-based directed (or forced) evolution studies have shown that it is sometimes possible to generate smaller enzymes that do retain activity.

It should be noted that although a large number of enzymes consist solely of protein, many also contain a non-protein component, known as a cofactor, that is necessary for the enzyme's catalytic activity. A cofactor may be another organic molecule, in which case it is called a coenzyme, or it may be an inorganic molecule, typically a metal ion such as iron, manganese, cobalt, copper or zinc. A coenzyme that binds tightly and permanently to the protein is generally referred to as the prosthetic group of the enzyme.

When an enzyme requires a cofactor for its activity, the inactive protein component is generally referred to as an apoenzyme, and the apoenzyme plus the cofactor (i.e. the active enzyme) is called a holoenzyme ( Figure 2 ).

An external file that holds a picture, illustration, etc.
Object name is bse0590001i002.jpg

The need for minerals and vitamins in the human diet is partly attributable to their roles within metabolism as cofactors and coenzymes.

Enzymes and reaction equilibrium

How do enzymes work? The broad answer to this question is that they do not alter the equilibrium (i.e. the thermodynamics) of a reaction. This is because enzymes do not fundamentally change the structure and energetics of the products and reagents, but rather they simply allow the reaction equilibrium to be attained more rapidly. Let us therefore begin by clarifying the concept of chemical equilibrium.

In many cases the equilibrium of a reaction is far ‘to the right’—that is, virtually all of the substrate (S) is converted into product (P). For this reason, reactions are often written as follows:

This is a simplification, as in all cases it is more correct to write this reaction as follows:

This indicates the presence of an equilibrium. To understand this concept it is perhaps most helpful to look at a reaction where the equilibrium point is quite central.

For example:

In this reaction, if we start with a solution of 1 mol l −1 glucose and add the enzyme, then upon completion we will have a mixture of approximately 0.5 mol l −1 glucose and 0.5 mol l −1 fructose. This is the equilibrium point of this particular reaction, and although it may only take a couple of seconds to reach this end point with the enzyme present, we would in fact come to the same point if we put glucose into solution and waited many months for the reaction to occur in the absence of the enzyme. Interestingly, we could also have started this reaction with a 1 mol l −1 fructose solution, and it would have proceeded in the opposite direction until the same equilibrium point had been reached.

The equilibrium point for this reaction is expressed by the equilibrium constant K eq as follows:

Thus for a reaction with central equilibrium, K eq = 1, for an equilibrium ‘to the right’ K eq is >1, and for an equilibrium ‘to the left’ K eq is <1.

Therefore if a reaction has a K eq value of 10 6 , the equilibrium is very far to the right and can be simplified by denoting it as a single arrow. We may often describe this type of reaction as ‘going to completion’. Conversely, if a reaction has a K eq value of 10 −6 , the equilibrium is very far to the left, and for all practical purposes it would not really be considered to proceed at all.

It should be noted that although the concentration of reactants has no effect on the equilibrium point, environmental factors such as pH and temperature can and do affect the position of the equilibrium.

It should also be noted that any biochemical reaction which occurs in vivo in a living system does not occur in isolation, but as part of a metabolic pathway, which makes it more difficult to conceptualize the relationship between reactants and reactions. In vivo reactions are not allowed to proceed to their equilibrium position. If they did, the reaction would essentially stop (i.e. the forward and reverse reactions would balance each other), and there would be no net flux through the pathway. However, in many complex biochemical pathways some of the individual reaction steps are close to equilibrium, whereas others are far from equilibrium, the latter (catalysed by regulatory enzymes) having the greatest capacity to control the overall flux of materials through the pathway.

Enzymes form complexes with their substrates

We often describe an enzyme-catalysed reaction as proceeding through three stages as follows:

The ES complex represents a position where the substrate (S) is bound to the enzyme (E) such that the reaction (whatever it might be) is made more favourable. As soon as the reaction has occurred, the product molecule (P) dissociates from the enzyme, which is then free to bind to another substrate molecule. At some point during this process the substrate is converted into an intermediate form (often called the transition state) and then into the product.

The exact mechanism whereby the enzyme acts to increase the rate of the reaction differs from one system to another. However, the general principle is that by binding of the substrate to the enzyme, the reaction involving the substrate is made more favourable by lowering the activation energy of the reaction.

In terms of energetics, reactions can be either exergonic (releasing energy) or endergonic (consuming energy). However, even in an exergonic reaction a small amount of energy, termed the activation energy, is needed to give the reaction a ‘kick start.’ A good analogy is that of a match, the head of which contains a mixture of energy-rich chemicals (phosphorus sesquisulfide and potassium chlorate). When a match burns it releases substantial amounts of light and heat energy (exergonically reacting with O 2 in the air). However, and perhaps fortunately, a match will not spontaneously ignite, but rather a small input of energy in the form of heat generated through friction (i.e. striking of the match) is needed to initiate the reaction. Of course once the match has been struck the amount of energy released is considerable, and greatly exceeds the small energy input during the striking process.

As shown in Figure 3 , enzymes are considered to lower the activation energy of a system by making it energetically easier for the transition state to form. In the presence of an enzyme catalyst, the formation of the transition state is energetically more favourable (i.e. it requires less energy for the ‘kick start’), thereby accelerating the rate at which the reaction will proceed, but not fundamentally changing the energy levels of either the reactant or the product.

An external file that holds a picture, illustration, etc.
Object name is bse0590001i003.jpg

Properties and mechanisms of enzyme action

Enzyme kinetics.

Enzyme kinetics is the study of factors that determine the speed of enzyme-catalysed reactions. It utilizes some mathematical equations that can be confusing to students when they first encounter them. However, the theory of kinetics is both logical and simple, and it is essential to develop an understanding of this subject in order to be able to appreciate the role of enzymes both in metabolism and in biotechnology.

Assays (measurements) of enzyme activity can be performed in either a discontinuous or continuous fashion. Discontinuous methods involve mixing the substrate and enzyme together and measuring the product formed after a set period of time, so these methods are generally easy and quick to perform. In general we would use such discontinuous assays when we know little about the system (and are making preliminary investigations), or alternatively when we know a great deal about the system and are certain that the time interval we are choosing is appropriate.

In continuous enzyme assays we would generally study the rate of an enzyme-catalysed reaction by mixing the enzyme with the substrate and continuously measuring the appearance of product over time. Of course we could equally well measure the rate of the reaction by measuring the disappearance of substrate over time. Apart from the actual direction (one increasing and one decreasing), the two values would be identical. In enzyme kinetics experiments, for convenience we very often use an artificial substrate called a chromogen that yields a brightly coloured product, making the reaction easy to follow using a colorimeter or a spectrophotometer. However, we could in fact use any available analytical equipment that has the capacity to measure the concentration of either the product or the substrate.

An external file that holds a picture, illustration, etc.
Object name is bse0590001in02.jpg

In almost all cases we would also add a buffer solution to the mixture. As we shall see, enzyme activity is strongly influenced by pH, so it is important to set the pH at a specific value and keep it constant throughout the experiment.

Our first enzyme kinetics experiment may therefore involve mixing a substrate solution (chromogen) with a buffer solution and adding the enzyme. This mixture would then be placed in a spectrophotometer and the appearance of the coloured product would be measured. This would enable us to follow a rapid reaction which, after a few seconds or minutes, might start to slow down, as shown in Figure 4 .

An external file that holds a picture, illustration, etc.
Object name is bse0590001i004.jpg

A common reason for this slowing down of the speed (rate) of the reaction is that the substrate within the mixture is being used up and thus becoming limiting. Alternatively, it may be that the enzyme is unstable and is denaturing over the course of the experiment, or it could be that the pH of the mixture is changing, as many reactions either consume or release protons. For these reasons, when we are asked to specify the rate of a reaction we do so early on, as soon as the enzyme has been added, and when none of the above-mentioned limitations apply. We refer to this initial rapid rate as the initial velocity ( v 0 ). Measurement of the reaction rate at this early stage is also quite straightforward, as the rate is effectively linear, so we can simply draw a straight line and measure the gradient (by dividing the concentration change by the time interval) in order to evaluate the reaction rate over this period.

We may now perform a range of similar enzyme assays to evaluate how the initial velocity changes when the substrate or enzyme concentration is altered, or when the pH is changed. These studies will help us to characterize the properties of the enzyme under study.

The relationship between enzyme concentration and the rate of the reaction is usually a simple one. If we repeat the experiment just described, but add 10% more enzyme, the reaction will be 10% faster, and if we double the enzyme concentration the reaction will proceed twice as fast. Thus there is a simple linear relationship between the reaction rate and the amount of enzyme available to catalyse the reaction ( Figure 5 ).

An external file that holds a picture, illustration, etc.
Object name is bse0590001i005.jpg

This relationship applies both to enzymes in vivo and to those used in biotechnological applications, where regulation of the amount of enzyme present may control reaction rates.

When we perform a series of enzyme assays using the same enzyme concentration, but with a range of different substrate concentrations, a slightly more complex relationship emerges, as shown in Figure 6 . Initially, when the substrate concentration is increased, the rate of reaction increases considerably. However, as the substrate concentration is increased further the effects on the reaction rate start to decline, until a stage is reached where increasing the substrate concentration has little further effect on the reaction rate. At this point the enzyme is considered to be coming close to saturation with substrate, and demonstrating its maximal velocity ( V max ). Note that this maximal velocity is in fact a theoretical limit that will not be truly achieved in any experiment, although we might come very close to it.

An external file that holds a picture, illustration, etc.
Object name is bse0590001i006.jpg

The relationship described here is a fairly common one, which a mathematician would immediately identify as a rectangular hyperbola. The equation that describes such a relationship is as follows:

The two constants a and b thus allow us to describe this hyperbolic relationship, just as with a linear relationship ( y = mx + c ), which can be expressed by the two constants m (the slope) and c (the intercept).

We have in fact already defined the constant a — it is V max . The constant b is a little more complex, as it is the value on the x -axis that gives half of the maximal value of y . In enzymology we refer to this as the Michaelis constant ( K m ), which is defined as the substrate concentration that gives half-maximal velocity.

Our final equation, usually called the Michaelis–Menten equation, therefore becomes:

In 1913, Leonor Michaelis and Maud Menten first showed that it was in fact possible to derive this equation mathematically from first principles, with some simple assumptions about the way in which an enzyme reacts with a substrate to form a product. Central to their derivation is the concept that the reaction takes place via the formation of an ES complex which, once formed, can either dissociate (productively) to release product, or else dissociate in the reverse direction without any formation of product. Thus the reaction can be represented as follows, with k 1 , k −1 and k 2 being the rate constants of the three individual reaction steps:

The Michaelis–Menten derivation requires two important assumptions. The first assumption is that we are considering the initial velocity of the reaction ( v 0 ), when the product concentration will be negligibly small (i.e. [S] ≫ [P]), such that we can ignore the possibility of any product reverting to substrate. The second assumption is that the concentration of substrate greatly exceeds the concentration of enzyme (i.e. [S]≫[E]).

The derivation begins with an equation for the expression of the initial rate, the rate of formation of product, as the rate at which the ES complex dissociates to form product. This is based upon the rate constant k 2 and the concentration of the ES complex, as follows:

Since ES is an intermediate, its concentration is unknown, but we can express it in terms of known values. In a steady-state approximation we can assume that although the concentration of substrate and product changes, the concentration of the ES complex itself remains constant. The rate of formation of the ES complex and the rate of its breakdown must therefore balance, where:

Hence, at steady state:

This equation can be rearranged to yield [ES] as follows:

The Michaelis constant K m can be defined as follows:

Equation 2 may thus be simplified to:

Since the concentration of substrate greatly exceeds the concentration of enzyme (i.e. [S] ≫ [E]), the concentration of uncombined substrate [S] is almost equal to the total concentration of substrate. The concentration of uncombined enzyme [E] is equal to the total enzyme concentration [E] T minus that combined with substrate [ES]. Introducing these terms to Equation 3 and solving for ES gives us the following:

We can then introduce this term into Equation 1 to give:

The term k 2 [E] T in fact represents V max , the maximal velocity. Thus Michaelis and Menten were able to derive their final equation as:

A more detailed derivation of the Michaelis–Menten equation can be found in many biochemistry textbooks (see section 4 of Recommended Reading section). There are also some very helpful web-based tutorials available on the subject.

Michaelis constants have been determined for many commonly used enzymes, and are typically in the lower millimolar range ( Table 5 ).

Enzyme (mmol l )
Carbonic anhydrase26
Chymotrypsin15
Ribonuclease8
Tyrosyl-tRNA synthetase0.9
Pepsin0.3

It should be noted that enzymes which catalyse the same reaction, but which are derived from different organisms, can have widely differing K m values. Furthermore, an enzyme with multiple substrates can have quite different K m values for each substrate.

A low K m value indicates that the enzyme requires only a small amount of substrate in order to become saturated. Therefore the maximum velocity is reached at relatively low substrate concentrations. A high K m value indicates the need for high substrate concentrations in order to achieve maximum reaction velocity. Thus we generally refer to K m as a measure of the affinity of the enzyme for its substrate—in fact it is an inverse measure, where a high K m indicates a low affinity, and vice versa.

The K m value tells us several important things about a particular enzyme.

  • An enzyme with a low K m value relative to the physiological concentration of substrate will probably always be saturated with substrate, and will therefore act at a constant rate, regardless of variations in the concentration of substrate within the physiological range.
  • An enzyme with a high K m value relative to the physiological concentration of substrate will not be saturated with substrate, and its activity will therefore vary according to the concentration of substrate, so the rate of formation of product will depend on the availability of substrate.
  • If an enzyme acts on several substrates, the substrate with the lowest K m value is frequently assumed to be that enzyme's ‘natural’ substrate, although this may not be true in all cases.
  • If two enzymes (with similar V max ) in different metabolic pathways compete for the same substrate, then if we know the K m values for the two enzymes we can predict the relative activity of the two pathways. Essentially the pathway that has the enzyme with the lower K m value is likely to be the ‘preferred pathway’, and more substrate will flow through that pathway under most conditions. For example, phosphofructokinase (PFK) is the enzyme that catalyses the first committed step in the glycolytic pathway, which generates energy in the form of ATP for the cell, whereas glucose-1-phosphate uridylyltransferase (GUT) is an enzyme early in the pathway leading to the synthesis of glycogen (an energy storage molecule). Both enzymes use hexose monophosphates as substrates, but the K m of PFK for its substrate is lower than that of GUT for its substrate. Thus at lower cellular hexose phosphate concentrations, PFK will be active and GUT will be largely inactive. At higher hexose phosphate concentrations both pathways will be active. This means that the cells only store glycogen in times of plenty, and always give preference to the pathway of ATP production, which is the more essential function.

Very often it is not possible to estimate K m values from a direct plot of velocity against substrate concentration (as shown in Figure 6 ) because we have not used high enough substrate concentrations to come even close to estimating maximal velocity, and therefore we cannot evaluate half-maximal velocity and thus K m . Fortunately, we can plot our experimental data in a slightly different way in order to obtain these values. The most commonly used alternative is the Lineweaver–Burk plot (often called the double-reciprocal plot). This plot linearizes the hyperbolic curved relationship, and the line produced is easy to extrapolate, allowing evaluation of V max and K m . For example, if we obtained only the first seven data points in Figure 6 , we would have difficulty estimating V max from a direct plot as shown in Figure 7 a.

An external file that holds a picture, illustration, etc.
Object name is bse0590001i007.jpg

However, as shown in Figure 7 b, if these seven points are plotted on a graph of 1/velocity against 1/substrate concentration (i.e. a double-reciprocal plot), the data are linearized, and the line can be easily extrapolated to the left to provide intercepts on both the y -axis and the x -axis, from which V max and K m , respectively, can be evaluated.

One significant practical drawback of using the Lineweaver–Burk plot is the excessive influence that it gives to measurements made at the lowest substrate concentrations. These concentrations might well be the most prone to error (due to difficulties in making multiple dilutions), and result in reaction rates that, because they are slow, might also be most prone to measurement error. Often, as shown in Figure 8 , such points when transformed on the Lineweaver–Burk plot have a significant impact on the line of best fit estimated from the data, and therefore on the extrapolated values of both V max and K m . The two sets of points shown in Figure 8 are identical except for the single point at the top right, which reflects (because of the plot's double-reciprocal nature) a single point derived from a very low substrate concentration and a low reaction rate. However, this single point can have an enormous impact on the line of best fit and the accompanying estimates of kinetic constants.

An external file that holds a picture, illustration, etc.
Object name is bse0590001i008.jpg

In fact there are other kinetic plots that can be used, including the Eadie–Hofstee plot, the Hanes plot and the Eisenthal–Cornish-Bowden plot, which are less prone to such problems. However, the Lineweaver–Burk plot is still the most commonly described kinetic plot in the majority of enzymology textbooks, and thus retains its influence in undergraduate education.

Enzymes are affected by pH and temperature

Various environmental factors are able to affect the rate of enzyme-catalysed reactions through reversible or irreversible changes in the protein structure. The effects of pH and temperature are generally well understood.

Most enzymes have a characteristic optimum pH at which the velocity of the catalysed reaction is maximal, and above and below which the velocity declines ( Figure 9 ).

An external file that holds a picture, illustration, etc.
Object name is bse0590001i009.jpg

The pH profile is dependent on a number of factors. As the pH changes, the ionization of groups both at the enzyme's active site and on the substrate can alter, influencing the rate of binding of the substrate to the active site. These effects are often reversible. For example, if we take an enzyme with an optimal pH (pH opt ) of 7.0 and place it in an environment at pH 6.0 or 8.0, the charge properties of the enzyme and the substrate may be suboptimal, such that binding and hence the reaction rate are lowered. If we then readjust the pH to 7.0, the optimal charge properties and hence the maximal activity of the enzyme are often restored. However, if we place the enzyme in a more extreme acidic or alkaline environment (e.g. at pH 1 or 14), although these conditions may not actually lead to changes in the very stable covalent structure of the protein (i.e. its configuration), they may well produce changes in the conformation (shape) of the protein such that, when it is returned to pH 7.0, the original conformation and hence the enzyme's full catalytic activity are not restored.

It should be noted that the optimum pH of an enzyme may not be identical to that of its normal intracellular surroundings. This indicates that the local pH can exert a controlling influence on enzyme activity.

The effects of temperature on enzyme activity are quite complex, and can be regarded as two forces acting simultaneously but in opposite directions. As the temperature is raised, the rate of molecular movement and hence the rate of reaction increases, but at the same time there is a progressive inactivation caused by denaturation of the enzyme protein. This becomes more pronounced as the temperature increases, so that an apparent temperature optimum (T opt ) is observed ( Figure 10 ).

An external file that holds a picture, illustration, etc.
Object name is bse0590001i010.jpg

Thermal denaturation is time dependent, and for an enzyme the term ‘optimum temperature’ has little real meaning unless the duration of exposure to that temperature is recorded. The thermal stability of an enzyme can be determined by first exposing the protein to a range of temperatures for a fixed period of time, and subsequently measuring its activity at one favourable temperature (e.g. 25°C).

The temperature at which denaturation becomes important varies from one enzyme to another. Normally it is negligible below 30°C, and starts to become appreciable above 40°C. Typically, enzymes derived from microbial sources show much higher thermal stability than do those from mammalian sources, and enzymes derived from extremely thermophilic microorganisms, such as thermolysin (a protease from Bacillus thermoproteolyticus ) and Taq polymerase (a DNA polymerase from Thermus aquaticus ), might be completely thermostable at 70°C and still retain substantial levels of activity even at 100°C.

Enzymes are sensitive to inhibitors

Substances that reduce the activity of an enzyme-catalysed reaction are known as inhibitors. They act by either directly or indirectly influencing the catalytic properties of the active site. Inhibitors can be foreign to the cell or natural components of it. Those in the latter category can represent an important element of the regulation of cell metabolism. Many toxins and also many pharmacologically active agents (both illegal drugs and prescription and over-the-counter medicines) act by inhibiting specific enzyme-catalysed processes.

Reversible inhibition

Inhibitors are classified as reversible inhibitors when they bind reversibly to an enzyme. A molecule that is structurally similar to the normal substrate may be able to bind reversibly to the enzyme's active site and therefore act as a competitive inhibitor. For example, malonate is a competitive inhibitor of the enzyme succinate dehydrogenase, as it is capable of binding to the enzyme's active site due to its close structural similarity to the enzyme's natural substrate, succinate (see below). When malonate occupies the active site of succinate dehydrogenase it prevents the natural substrate, succinate, from binding, thereby slowing down the rate of oxidation of succinate to fumarate (i.e. inhibiting the reaction).

An external file that holds a picture, illustration, etc.
Object name is bse0590001in03.jpg

One of the characteristics of competitive inhibitors is that they can be displaced from the active site if high concentrations of substrate are used, thereby restoring enzyme activity. Thus competitive inhibitors increase the K m of a reaction because they increase the concentration of substrate required to saturate the enzyme. However, they do not change V max itself.

In the case of certain enzymes, high concentrations of either the substrate or the product can be inhibitory. For example, invertase activity is considerably reduced in the presence of high concentrations of sucrose (its substrate), whereas the β-galactosidase of Aspergillus niger is strongly inhibited by galactose (its product). Products of an enzyme reaction are some of the most commonly encountered competitive inhibitors.

Other types of reversible inhibitor also exist. Non-competitive inhibitors react with the enzyme at a site distinct from the active site. Therefore the binding of the inhibitor does not physically block the substrate–binding site, but it does prevent subsequent reaction. Most non-competitive inhibitors are chemically unrelated to the substrate, and their inhibition cannot be overcome by increasing the substrate concentration. Such inhibitors in effect reduce the concentration of the active enzyme in solution, thereby reducing the V max of the reaction. However, they do not change the value of K m .

Uncompetitive inhibition is rather rare, occurring when the inhibitor is only able to bind to the enzyme once a substrate molecule has itself bound. As such, inhibition is most significant at high substrate concentrations, and results in a reduction in the V max of the reaction. Uncompetitive inhibition also causes a reduction in K m , which seems somewhat counterintuitive as this means that the affinity of the enzyme for its substrate is actually increased when the inhibitor is present. This effect occurs because the binding of the inhibitor to the ES complex effectively removes ES complex and thereby affects the overall equilibrium of the reaction favouring ES complex formation. It is noteworthy however that since both V max and K m are reduced the observed reaction rates with inhibitor present are always lower than those in the absence of the uncompetitive inhibitor.

Irreversible inhibitors and poisons

If an inhibitor binds permanently to an enzyme it is known as an irreversible inhibitor. Many irreversible inhibitors are therefore potent toxins.

Organophosphorus compounds such as diisopropyl fluorophosphate (DFP) inhibit acetylcholinesterase activity by reacting covalently with an important serine residue found within the active site of the enzyme. The physiological effect of this inactivation is interference with neurotransmitter inactivation at the synapses of nerves, resulting in the constant propagation of nerve impulses, which can lead to death. DFP was originally evaluated by the British as a chemical warfare agent during World War Two, and modified versions of this compound are now widely used as organophosphate pesticides (e.g. parathione, malathione).

Allosteric regulators and the control of enzyme activity

Having spent time learning about enzyme kinetics and the Michaelis–Menten relationship, it is often quite disconcerting to find that some of the most important enzymes do not in fact display such properties. Allosteric enzymes are key regulatory enzymes that control the activities of metabolic pathways by responding to inhibitors and activators. These enzymes in fact show a sigmoidal (S-shaped) relationship between reaction rate and substrate concentration ( Figure 11 ), rather than the usual hyperbolic relationship. Thus for allosteric enzymes there is an area where activity is lower than that of an equivalent ‘normal’ enzyme, and also an area where activity is higher than that of an equivalent ‘normal’ enzyme, with a rapid transition between these two phases. This is rather like a switch that can quickly be changed from ‘off’ (low activity) to ‘on’ (full activity).

An external file that holds a picture, illustration, etc.
Object name is bse0590001i011.jpg

Most allosteric enzymes are polymeric—that is, they are composed of at least two (and often many more) individual polypeptide chains. They also have multiple active sites where the substrate can bind. Much of our understanding of the function of allosteric enzymes comes from studies of haemoglobin which, although it is not an enzyme, binds oxygen in a similarly co-operative way and thus also demonstrates this sigmoidal relationship. Allosteric enzymes have an initially low affinity for the substrate, but when a single substrate molecule binds, this may break some bonds within the enzyme and thereby change the shape of the protein such that the remaining active sites are able to bind with a higher affinity. Therefore allosteric enzymes are often described as moving from a tensed state or T-state (low affinity) in which no substrate is bound, to a relaxed state or R-state (high affinity) as substrate binds. Other molecules can also bind to allosteric enzymes, at additional regulatory sites (i.e. not at the active site). Molecules that stabilize the protein in its T-state therefore act as allosteric inhibitors, whereas molecules that move the protein to its R-state will act as allosteric activators or promoters.

A good example of an allosteric enzyme is aspartate transcarbamoylase (ATCase), a key regulatory enzyme that catalyses the first committed step in the sequence of reactions that produce the pyrimidine nucleotides which are essential components of DNA and RNA. The reaction is as follows:

An external file that holds a picture, illustration, etc.
Object name is bse0590001in04.jpg

The end product in the pathway, the pyrimidine nucleotide cytidine triphosphate (CTP), is an active allosteric inhibitor of the enzyme ATCase. Therefore when there is a high concentration of CTP in the cell, this feeds back and inhibits the ATCase enzyme, reducing its activity and thus lowering the rate of production of further pyrimidine nucleotides. As the concentration of CTP in the cell decreases then so does the inhibition of ATCase, and the resulting increase in enzyme activity leads to the production of more pyrimidine nucleotides. This negative feedback inhibition is an important element of biochemical homeostasis within the cell. However, in order to synthesize DNA and RNA, the cell requires not only pyrimidine nucleotides but also purine nucleotides, and these are needed in roughly equal proportions. Purine synthesis occurs through a different pathway, but interestingly the final product, the purine nucleotide adenosine triphosphate (ATP), is a potent activator of the enzyme ATCase. This is logical, since when the cell contains high concentrations of purine nucleotides it will require equally high concentrations of pyrimidine nucleotides in order for these two types of nucleotide to combine to form the polymers DNA and RNA. Thus ATCase is able to regulate the production of pyrimidine nucleotides within the cell according to cellular demand, and also to ensure that pyrimidine nucleotide synthesis is synchronized with purine nucleotide synthesis—an elegant biochemical mechanism for the regulation of an extremely important metabolic process.

There are some rare, although important, cases of monomeric enzymes that have only one substrate-binding site but are capable of demonstrating the sigmoidal reaction kinetics characteristic of allosteric enzymes. Particularly noteworthy in this context is the monomeric enzyme glucokinase (also called hexokinase IV), which catalyses the phosphorylation of glucose to glucose-6-phosphate (which may then either be metabolized by the glycolytic pathway or be used in glycogen synthesis). It has been postulated that this kinetic behaviour is a result of individual glucokinase molecules existing in one of two forms—a low-affinity form and a high-affinity form. The low-affinity form of the enzyme reacts with its substrate (glucose), is then turned into the high-affinity form, and remains in that state for a short time before slowly returning to its original low-affinity form (demonstrating a so-called slow transition). Therefore at high substrate concentrations the enzyme is likely to react with a second substrate molecule soon after the first one (i.e. while still in its high-affinity form), whereas at lower substrate concentrations the enzyme may transition back to its low-affinity form before it reacts with subsequent substrate molecules. This results in its characteristic sigmoidal reaction kinetics.

Origin, purification and uses of enzymes

Enzymes are ubiquitous.

Enzymes are essential components of animals, plants and microorganisms, due to the fact that they catalyse and co-ordinate the complex reactions of cellular metabolism.

Up until the 1970s, most of the commercial application of enzymes involved animal and plant sources. At that time, bulk enzymes were generally only used within the food-processing industry, and enzymes from animals and plants were preferred, as they were considered to be free from the problems of toxicity and contamination that were associated with enzymes of microbial origin. However, as demand grew and as fermentation technology developed, the competitive cost of microbial enzymes was recognized and they became more widely used.

Compared with enzymes from plant and animal sources, microbial enzymes have economic, technical and ethical advantages, which will now be outlined.

Economic advantages

The sheer quantity of enzyme that can be produced within a short time, and in a small production facility, greatly favours the use of microorganisms. For example, during the production of rennin (a milk-coagulating enzyme used in cheese manufacture) the traditional approach is to use the enzyme extracted from the stomach of a calf (a young cow still feeding on its mother's milk). The average quantity of rennet extracted from a calf's stomach is 10 kg, and it takes several months of intensive farming to produce a calf. In comparison, a 1 000-litre fermenter of recombinant Bacillus subtilis can produce 20 kg of enzyme within 12 h. Thus the microbial product is clearly preferable economically, and is free from the ethical issues that surround the use of animals. Indeed, most of the cheese now sold in supermarkets is made from milk coagulated with microbial enzymes (so is suitable for vegetarians).

A further advantage of using microbial enzymes is their ease of extraction. Many of the microbial enzymes used in biotechnological processes are secreted extracellularly, which greatly simplifies their extraction and purification. Microbial intracellular enzymes are also often easier to obtain than the equivalent animal or plant enzymes, as they generally require fewer extraction and purification steps.

Animal and plant sources usually need to be transported to the extraction facility, whereas when microorganisms are used the same facility can generally be employed for production and extraction. In addition, commercially important animal and plant enzymes are often located within only one organ or tissue, so the remaining material is essentially a waste product, disposal of which is required.

Finally, enzymes from plant and animal sources show wide variation in yield, and may only be available at certain times of year, whereas none of these problems are associated with microbial enzymes.

Technical advantages

Microbial enzymes often have properties that make them more suitable for commercial exploitation. In comparison with enzymes from animal and plant sources, the stability of microbial enzymes is usually high. For example, the high temperature stability of enzymes from thermophilic microorganisms is often useful when the process must operate at high temperatures (e.g. during starch processing).

Microorganisms are also very amenable to genetic modification to produce novel or altered enzymes, using relatively simple methods such as plasmid insertion. The genetic manipulation of animals and plants is technically much more difficult, is more expensive and is still the subject of significant ethical concern, especially in the U.K.

Enzymes may be intracellular or extracellular

Although many enzymes are retained within the cell, and may be located in specific subcellular compartments, others are released into the surrounding environment. The majority of enzymes in industrial use are extracellular proteins from either fungal sources (e.g. Aspergillus species) or bacterial sources (e.g. Bacillus species). Examples of these include α-amylase, cellulase, dextranase, proteases and amyloglucosidase. Many other enzymes for non-industrial use are intracellular and are produced in much smaller amounts by the cell. Examples of these include asparaginase, catalase, cholesterol oxidase, glucose oxidase and glucose-6-phosphate dehydrogenase.

Enzyme purification

Within the cell, enzymes are generally found along with other proteins, nucleic acids, polysaccharides and lipids. The activity of the enzyme in relation to the total protein present (i.e. the specific activity) can be determined and used as a measure of enzyme purity. A variety of methods can be used to remove contaminating material in order to purify the enzyme and increase its specific activity. Enzymes that are used as diagnostic reagents and in clinical therapeutics are normally prepared to a high degree of purity, because great emphasis is placed on the specificity of the reaction that is being catalysed. Clearly the higher the level of purification, the greater the cost of enzyme production. In the case of many bulk industrial enzymes the degree of purification is less important, and such enzymes may often be sold as very crude preparations of culture broth containing the growth medium, organisms (whole or fragmented) and enzymes of interest. However, even when the cheapest bulk enzymes are utilized (e.g. proteases for use in washing powders), the enzyme cost can contribute around 5–10% of the final product value.

Pretreatment

At the end of a fermentation in which a microorganism rich in the required enzyme has been cultured, the broth may be cooled rapidly to 5°C to prevent further microbial growth and stabilize the enzyme product. The pH may also be adjusted to optimize enzyme stability. If the enzyme-producing organism is a fungus, this may be removed by centrifugation at low speed. If the enzyme source is bacterial, the bacteria are often flocculated with aluminum sulfate or calcium chloride, which negate the charge on the bacterial membranes, causing them to clump and thus come out of suspension.

Extracellular enzymes are found in the liquid component of the pretreatment process. However, intracellular enzymes require more extensive treatment. The biomass may be concentrated by centrifugation and washed to remove medium components. The cellular component must then be ruptured to release the enzyme content. This can be done using one or more of the following processes:

  • • ball milling (using glass beads)
  • • enzymic removal of the cell wall
  • • freeze–thaw cycles
  • • liquid shearing through a small orifice at high pressure (e.g. within a French press)
  • • osmotic shock
  • • sonication.

Separation of enzymes from the resulting solution may then involve a variety of separation processes, which are often employed in a sequential fashion.

The first step in an enzyme purification procedure commonly involves separation of the proteins from the non-protein components by a process of salting out. Proteins remain in aqueous solution because of interactions between the hydrophilic (water-loving) amino acids and the surrounding water molecules (the solvent). If the ionic strength of the solvent is increased by adding an agent such as ammonium sulfate, some of the water molecules will interact with the salt ions, thereby decreasing the number of water molecules available to interact with the protein. Under such conditions, when protein molecules cannot interact with the solvent, they interact with each other, coagulating and coming out of solution in the form of a precipitate. This precipitate (containing the enzyme of interest and other proteins) can then be filtered or centrifuged, and separated from the supernatant.

Since different proteins vary in the extent to which they interact with water, it is possible to perform this process using a series of additions of ammonium sulfate, increasing the ionic strength in a stepwise fashion and removing the precipitate at each stage. Thus such fractional precipitation is not only capable of separating protein from non-protein components, but can also enable separation of the enzyme of interest from some of the other protein components.

Subsequently a wide variety of techniques may be used for further purification, and steps involving chromatography are standard practice.

Ion-exchange chromatography is often effective during the early stages of the purification process. The protein solution is added to a column containing an insoluble polymer (e.g. cellulose) that has been modified so that its ionic characteristics will determine the type of mobile ion (i.e. cation or anion) it attracts. Proteins whose net charge is opposite to that of the ion-exchange material will bind to it, whereas all other proteins will pass through the column. A subsequent change in pH or the introduction of a salt solution will alter the electrostatic forces, allowing the retained protein to be released into solution again.

Gel filtration can be utilized in the later stages of a purification protocol to separate molecules on the basis of molecular size. Columns containing a bed of cross-linked gel particles such as Sephadex are used. These gel particles exclude large protein molecules while allowing the entry of smaller molecules. Separation occurs because the larger protein molecules follow a path down the column between the Sephadex particles (occupying a smaller fraction of the column volume). Larger molecules therefore have a shorter elution time and are recovered first from the gel filtration column.

Affinity chromatography procedures can often enable purification protocols to be substantially simplified. Typically, with respect to enzyme purification, a column would be packed with a particulate stationary phase to which a ligand molecule such as a substrate analogue, inhibitor or cofactor of the enzyme of interest would be firmly bound. As the sample mixture is passed through the column, the enzyme interacts with, and binds, to the immobilised ligand, being retained within the column as all of the other components of the mixture pass through the column unrewarded. Subsequently a solution of the ligand is introduced to the column to release (elute) and thereby recover the bound enzyme from the column in a highly purified form.

Nowadays numerous alternative affinity chromatography procedures exist that are able to separate enzymes by binding to areas of the molecule away form their active site. Advances in molecular biology enable us to purify recombinant proteins, including enzymes, through affinity tagging. In a typical approach the gene for the enzyme of interest would be modified to code for a further short amino acid sequence at either the N- or C- terminal. For example, a range of polyhistidine tagging procedures are available to yield protein products with six or more consecutive histidine residues at their N- or C- terminal end. When a mixture containing the tagged protein of interest is subsequently passed through a column containing a nickel-nitrilotriacetic acid (Ni-NTA) agarose resin, the histidine residues on the recombinant protein bind to the nickel ions attached to the support resin, retaining the protein, whilst other protein and non-protein components pass through the column. Elution of the bound protein can then be accomplished by adding imidazole to the column, or by reducing the pH to 5-6 to displace the His-tagged protein from the nickel ions.

Such techniques are therefore capable of rapidly and highly effectively isolating an enzyme from a complex mixture in only one step, and typically provide protein purities of up to 95%. If more highly purified enzyme products are required, other supplemental options are also available, including various forms of preparative electrophoresis e.g. disc-gel electrophoresis and isoelectric focusing.

Finishing of enzymes

Enzymes are antigenic, and since problems occurred in the late 1960s when manufacturing workers exhibited severe allergic responses after breathing enzyme dusts, procedures have now been implemented to reduce dust formation. These involve supplying enzymes as liquids wherever possible, or increasing the particle size of dry powders from 10 μm to 200–500 μm by either prilling (mixing the enzyme with polyethylene glycol and preparing small spheres by atomization) or marumerizing (mixing the enzyme with a binder and water, extruding long filaments, converting them into spheres in a marumerizer, drying them and covering them with a waxy coat).

Industrial enzymology

Although many industrial processes, such as cheese manufacturing, have traditionally used impure enzyme sources, often from animals or plants, the development of much of modern industrial enzymology has gone hand in hand with the commercial exploitation of microbial enzymes. These were introduced to the West in around 1890 when the Japanese scientist Jokichi Takamine settled in the U.S.A. and set up an enzyme factory based on Japanese technology. The principal product was Takadiastase, a mixture of amylolytic and proteolytic enzymes prepared by cultivating the fungus Aspergillus oryzae on rice or wheat bran. Takadiastase was marketed successfully in the U.S.A. as a digestive aid for the treatment of dyspepsia, which was then believed to result from the incomplete digestion of starch.

Bacterial enzymes were developed in France by August Boidin and Jean Effront, who in 1913 found that Bacillus subtilis produced a heat-stable α-amylase when grown in a liquid medium made by extraction of malt or grain. The enzyme was primarily used within the textile industry for the removal of the starch that protects the warp in the manufacture of cotton.

In around 1930 it was found that fungal pectinases could be used in the preparation of fruit products. In subsequent years, several other hydrolases were developed and sold commercially (e.g. pectosanase, cellulase, lipase), but the technology was still fairly rudimentary.

After World War Two the fermentation industry underwent rapid development as methods for the production of antibiotics were developed. These methods were soon adapted for the production of enzymes. In the 1960s, glucoamylase was introduced as a means of hydrolysing starch, replacing acid hydrolysis. Subsequently, in the 1960s and 1970s, proteases were incorporated into detergents and then glucose isomerase was introduced to produce sweetening agents in the form of high-fructose syrups. Since the 1990s, lipases have been incorporated into washing powders, and a variety of immobilized enzyme processes have been developed (see section on enzyme immobilization), many of which utilize intracellular enzymes.

Currently, enzymes are used in four distinct fields of commerce and technology ( Table 6 ):

  • • as industrial catalysts
  • • as therapeutic agents
  • • as analytic reagents
  • • as manipulative tools (e.g. in genetics).
EnzymeReactionSourceApplication
Acid proteasesProtein digestion , Milk coagulation in cheese manufacture
Alkaline proteasesProtein digestion speciesDetergents and washing powders
AminoacylaseHydrolysis of acylated l–amino acids speciesProduction of l–amino acids
α-AmylaseStarch hydrolysis speciesConversion of starch to glucose or dextrans in the food industry
AmyloglucosidaseDextrin hydrolysis speciesGlucose production
β-GalactosidaseLactose hydrolysis speciesHydrolysis of lactose in milk or whey
Glucose isomeraseConversion of glucose to fructose speciesHigh-fructose syrup production
Penicillin acylasePenicillin side-chain cleavage 6-APA formation for production of semi-synthetic penicillins
l-AsparaginaseRemoval of l–asparagine essential for tumour growth Cancer chemotherapy, particularly for leukaemia
UrokinasePlasminogen activationHumanRemoval of fibrin clots from bloodstream
Glucose oxidaseGlucose oxidation Detection of glucose in blood
LuciferaseBioluminescenceMarine bacteria or fireflyBioluminescent assays involving ATP
PeroxidaseDye oxidation using H O HorseradishQuantification of hormones and antibodies
UreaseHydrolysis of urea to CO and NH Jack beanUrea quantification in body fluids
LysozymeHydrolysis of 1–4 glycosidic bondsHen egg whiteDisruption of mucopeptide in bacterial cell walls
NucleasesHydrolysis of phosphodiester bondsVarious bacteriaRestriction enzymes used in genetic manipulation to cut DNA
DNA polymerasesDNA synthesis DNA amplification used in the polymerase chain reaction

Of the thousands of different types of enzymes, about 95% are available from suppliers in quantities ranging from μg to kg, provided essentially for research purposes. Around 40–50 enzymes are produced on an industrial scale (i.e. ranging from multiple kilograms to tonnes per annum). The global enzyme market is currently dominated by the hydrolases, especially the proteases, together with amylases, cellulases and lipases supplied either as liquid concentrates or as powders or granules that release the soluble enzyme on dissolution. Global production is dominated by two companies, which between them supply more than two-thirds of the global enzyme market, namely the Danish company Novozymes, with a market share of 47%, and the U.S. company DuPont (which recently acquired Genencor), with 21%.

The value of the world enzyme market has increased steadily from £110 million in 1960 to £200 million in 1970, £270 million in 1980, £1 000 million in 1990 and over £2 000 million in 2010. Food and beverage enzymes represented the largest sector of the industrial enzymes market in 2010, with a value of £750 million, and the market for enzymes for technical applications (including diagnostic applications, research and biotechnology) accounted for a further £700 million. Estimates of future demand are in the range of £4 000–5 000 million between 2015 and 2016, growing at a rate of 6–7% annually. The developing economies of the Asia-Pacific Region, the Middle East and Africa are now seen to be emerging as the fastest growing markets for industrial enzymes.

Microbial enzymes are typically produced in batches by culturing the producing organism within a batch fermenter. Fermentation typically lasts between 30 and 150 h, with the optimum enzyme yield for the process falling somewhere between the optimum biomass yield and the point of maximal enzyme activity within the cells. Relatively small fermenters with a volume of 10–100 m 3 are generally employed, allowing flexibility where a number of different products are being produced. Many production systems are optimized by means of a fed-batch process, in which substrates are gradually fed into the reactor over the course of the fermentation, rather than being provided all at once at the start of the process. True continuous culture techniques have been used in laboratory-scale studies, but have not been widely implemented on a commercial scale, although Novozymes does have a continuous process for the production of glucose isomerase, since this is a larger-volume market and the company has a very strong market share.

Enzyme immobilization

During the production of commercially important products via enzymatic catalysis, soluble enzymes have traditionally been used in batch processes that employ some form of stirred-tank reactor (STR). In these processes, at the end of the batch run the product must be separated from any unused substrate, and also from the enzyme catalyst. Removal of the enzyme at this stage can be achieved by thermal denaturation (only if the product is thermostable) or by ammonium sulfate precipitation or ultrafiltration. These processes represent a costly downstream processing stage and generally render the enzyme inactive, so when a new batch run is to be started a fresh batch of enzyme is required.

Immobilized enzyme systems, in contrast, ‘fix’ the enzyme so that it can be reused many times, which has a significant impact on production costs. As a very simple example, if an enzyme is mixed with a solution of warm (but not too hot) agar and this is allowed to set, the enzyme will be entrapped (for the purposes of this example let us ignore the fact that the enzyme will gradually leak out of this gel). The agar can then be cut up into cubes and these can be placed in a STR, together with substrate, as shown in Figure 12 . Again the reaction would be allowed to proceed (and it might actually be slower due to diffusional constraints and other effects described later). At the end of the batch run the catalyst can now be easily separated from the product by passing the reactor contents through a coarse mesh. Immediately an important downstream processing step has been carried out and, just as importantly, the active enzyme has been recovered so that it can be reused for the next batch run. This ease of separation of enzyme from product is a major advantage of all immobilized systems over their counterparts that use free (i.e. soluble) enzyme.

An external file that holds a picture, illustration, etc.
Object name is bse0590001i012.jpg

This physical advantage of ease of reuse of immobilized biocatalysts is one of the main reasons why such systems are favoured commercially. However, immobilization may also produce biochemical changes that lead to enhanced biocatalyst stability, which may be manifested as:

  • • an increased rate of catalysis
  • • prolonged duration of catalysis
  • • greater operational stability to extremes of pH, temperature, etc.

The particular advantage(s) conferred by immobilization will therefore differ from one system to another. It should be noted that often there may be no biochemical advantage at all, and the simple physical advantage of ease of separation of the biocatalyst from the product may be sufficient to favour the commercial development of an immobilized process.

At this point one problem that will immediately spring to mind for most students is that they have always been taught to fully mix all of the reagents of a reaction, yet the basic principle of immobilization is to partition the biocatalyst into a distinct phase, rather than mix it homogeneously with the substrate. Will this not cause reaction rates to be low? The answer to this question is yes, and the relationship between the activity of an immobilized system and a non-immobilized system can be expressed as the effectiveness factor (η), where:

Thus an immobilized system with an effectiveness factor of 0.1 would show only 10% of the activity of a non-immobilized system with the same amount of enzyme and operating under the same conditions. At first sight this might appear to be a major problem. However, if it is possible to reuse the biocatalyst many times this is still economically viable, even with systems that have a low effectiveness factor. In principle, therefore, for economic viability:

Thus if an immobilized system has an effectiveness factor of 0.1 (i.e. 10%) and we can reuse the biocatalyst 10 times, we essentially achieve the same overall catalytic activity with both the non-immobilized system and the immobilized one. However, if we are able to reuse the biocatalyst 100 times we in fact obtain 10 times more total activity from the immobilized system than from the equivalent non-immobilized system, so the immobilized system may be economically preferable.

Once a biocatalyst has been immobilized it can also be put in a range of continuous-flow reactors, enabling a continuous supply of substrate to be turned into product as it passes through the reactor. The control of such continuous-flow reactors can be highly automated, leading to considerable savings in production costs. For example, a STR can be easily modified to produce a continuous-flow stirred-tank reactor (CSTR) ( Figure 13 a), in which the enzyme is held within the reactor by a coarse mesh, and the product continuously flows out of the reactor as substrate is pumped in. It is also possible to produce a packed-bed reactor (PBR) ( Figure 13 b), in which the agar cubes are packed into a column and the substrate is pumped through the bed without any need for stirring.

An external file that holds a picture, illustration, etc.
Object name is bse0590001i013.jpg

CSTRs and PBRs enable the enzyme to be reused many times before it needs to be replaced. For example, in the production of high-fructose syrups, the immobilized glucose isomerase enzyme would typically be used continuously for between 2 and 4 months, and only after this time (when its activity would have dropped to 25% of the original level) would it need to be replaced.

The overall operating costs of continuous-flow reactors are often significantly lower than those of equivalent batch processes. Batch reactors need to be emptied and refilled frequently at regular intervals. Not only is this procedure expensive, but it also means that there are considerable periods of time when such reactors are not productive (so-called ‘downtime’). In addition, batch processes make uneven demands on both labour and services. They may also result in pronounced batch-to-batch variations, as the reaction conditions change with time, and they may be difficult to scale up, due to the changing power requirements for efficient mixing. Due to their higher overall process efficiency, continuous processes using immobilized enzymes may be undertaken in production facilities that are around 10 to 100 times smaller than those required for equivalent batch processes using soluble enzymes. Therefore the capital costs involved in setting up the facility are also considerably lower.

Immobilization techniques

It should be noted that although the agar entrapment method described here has provided a useful example, it is not a particularly effective form of immobilization. The high temperature required to prevent the agar from setting may lead to thermal inactivation of the enzyme, and the agar gel itself is very porous and will allow the enzyme to leak out into the surrounding solution.

There are in fact thousands of different techniques of immobilization, all of which are much more effective than our example. In general these techniques can be classified as belonging to one of three categories ( Figure 14 ):

  • • adsorption
  • • covalent bonding
  • • entrapment.

An external file that holds a picture, illustration, etc.
Object name is bse0590001i014.jpg

The physical adsorption of an enzyme to a supporting matrix is the oldest method of immobilization. As early as 1916, J.M. Nelson and Edward G. Griffin described the adsorption of yeast invertase on to activated charcoal, and the subsequent use of this preparation for sucrose hydrolysis. Over the years a variety of adsorbents have been used, including cellulose, Sephadex, polystyrene, kaolinite, collagen, alumina, silica gel and glass. Such immobilization procedures are extremely easy to perform, as the adsorbent and enzyme are simply stirred together for a time (typically minutes to hours). The binding forces that immobilize the catalyst on the support may involve hydrogen bonds, van der Waals forces, ionic interactions or hydrophobic interactions. Such forces are generally weak in comparison with covalent bonds—for example, a hydrogen bond has an energy content of about 20 kJ mol −1 , compared with 200–500 kJ mol −1 for a covalent bond. Thus, when using such methods, yields (i.e. the amount of enzyme bound per unit of adsorbent) are generally low. In addition, adsorption is generally easily reversed, and can lead to desorption of the enzyme at a critical time.

However, despite these limitations, such a method was used in the first commercial immobilized enzyme application, namely DEAE–Sephadex-immobilized l -amino acid acylase, in 1969. DEAE–Sephadex is an ion-exchange resin that consists of an inert dextran particle activated by the addition of numerous diethylaminoethyl groups. Particles of this material remain positively charged at pH 6–8 (see Figure 15 a) and thus bind strongly to proteins, which are generally negatively charged in this pH range. If the pH is kept constant, the enzyme and support will remain ionically linked. However, when over time the enzyme loses its activity through denaturation, the pH can be adjusted to a more acidic value, the old enzyme will be desorbed, and the pH can then be readjusted back to pH 6–8 and a fresh batch of enzyme bound. Thus the support matrix may be used many times, giving the process significant economic benefits.

An external file that holds a picture, illustration, etc.
Object name is bse0590001i015.jpg

Clearly DEAE–Sephadex immobilization is only of value for enzymes that have a neutral-to-alkaline pH optimum. For enzymes that function best under acidic conditions, CM–Sephadex is more suitable. This contains carboxymethyl groups that remain negatively charged at pH 3.5–4.5 ( Figure 15 b). Proteins at this pH are generally positively charged and will thus ionically bind to the support. Desorption of the enzyme will occur when the pH is adjusted to a more alkaline value.

Due to the simplicity and controllability of this immobilization procedure, combined with the economic benefits of reuse of the support, ion-exchange materials are now widely used as the method of choice in many industrial settings.

Covalent bonding

Immobilization of enzymes by covalent bonding to activated polymers is a widely used approach since, although it is often a tedious procedure, it is capable of producing an immobilized enzyme that is firmly bound to its support. The range of polymers and chemical coupling procedures that are used is enormous.

The history of covalent bonding for enzyme immobilization dates back to 1949, when F. Michael and J. Ewers used the azide derivative of carboxymethylcellulose to immobilize a variety of proteins. Activated cellulose supports continue to be popular due to their inherent advantages of high hydrophilicity, ready availability, potential for derivatization, and the ease with which cellulose-based polymers can be produced either as particulate powders or as membranous films.

It is often more effective not to build the reactive group into the cellulose itself, but instead to use a chemical ‘bridge’ between the cellulose and the enzyme molecule. The requirements for such a bridging or linking molecule are that it must be small, and that once it has reacted with the support it must have a further reactive group capable of reacting with the enzyme. An example of such a bridging molecule is glutaraldehyde, which contains two aldehyde groups, one at either end of its (CH 2 ) 3 moiety. At neutral pH values the aldehyde groups will react with free amino groups. Thus one end of the glutaraldehyde molecule may be attached to the support, and the other to the enzyme.

Covalently immobilized enzymes are strongly bound to their support, so when the proteins denature they are difficult to remove (in contrast to adsorption, as described earlier). Therefore it is usual for both the enzyme and the support to be replaced. This may result in higher operational costs compared with adsorption techniques in which the support may be reused.

The entrapment of an enzyme can be achieved in a number of ways:

  • • inclusion within the matrix of a highly cross-linked polymer
  • • separation from the bulk phase by a semi-permeable ‘microcapsule’
  • • dissolution in a distinct non-aqueous phase.

An important feature of entrapment techniques is that the enzyme is not in fact attached to anything. Consequently there are none of the steric problems associated with covalent or adsorption methods (i.e. the possibility of the enzyme binding in such a way that its active site is obstructed by part of the supporting polymer matrix).

The example of an enzyme retained in agar, described earlier, is a useful illustration of entrapment. A preferable alternative involves mixing the catalyst with sodium alginate gel and extruding this into a solution of calcium chloride to produce solid calcium alginate particles. This technique has the advantage of not requiring the use of high temperatures. However, although it is a popular activity in teaching laboratories, outside that setting it is generally unsuitable for the immobilization of purified enzymes, as these are often able to leak out of the gel. Entrapment techniques for purified enzymes are more likely to involve retaining the enzyme behind some form of ultrafiltration membrane. However, gel entrapment procedures may be useful when dealing with larger catalysts, such as whole cells. For example, gel-immobilized living yeast cells have been used successfully in the manufacture of champagne by Moët & Chandon.

Immobilization: changes in enzyme properties

Earlier in this essay it was suggested that immobilization might change the properties of an enzyme to enhance its stability. Initially it was believed that such enhanced stability resulted from the formation of bonds between the enzyme and the supporting matrix that physically stabilize the structure of the protein. Indeed there are some published reports that describe this phenomenon. With regard to the stabilization of proteolytic enzymes, which often exhibit more prolonged activity in the immobilized state, this is most probably explained by the fact that such proteases in free solution are prone to autodigestion (i.e. enzyme molecules cleave the peptide bonds of adjacent enzyme molecules), a process that is largely prevented when they are fixed to a supporting matrix.

However, the effects of immobilization are more often due to the supporting matrix changing the microenvironment around the enzyme and/or introducing diffusional constraints that modify the activity of the catalyst. Consider, for example, immobilization of the enzyme by adsorption on to a polyanionic (negatively charged) support such as cellulose. If the substrate is a cation (i.e. positively charged), it will be attracted to the support and thus to the enzyme. In this case the enzyme might well display higher activity, as the substrate concentration in its microenvironment would be higher than that in the surrounding bulk phase. Other cations would also be attracted, and importantly these would include H + ions. Thus the microenvironment would also be enriched in H + ions, so the pH surrounding the enzyme would be lower than the pH of the bulk phase. Consequently the enzyme would also exhibit an altered pH profile compared with that of its soluble counterpart.

In addition, the immobilization matrix might act as a barrier to the diffusion of substrates, products and other molecules. For example, if a high enzyme loading was put into a gel particle and this was then immersed in substrate solution, the substrate would diffuse into the gel and rapidly be converted into product. Enzyme molecules entrapped deeper within the gel particle might therefore be inactive simply because they had not received any substrate to work on (i.e. all of the substrate was converted to product in the outer layers of the particle). Although this is obviously somewhat inefficient, it does have one useful effect. When over time the enzyme within the system denatures, the loss of activity of the enzyme in the outer part of the particle means that substrate will now diffuse deeper into the particle to reach the previously unused core enzyme molecules. In effect this inner reserve of enzyme will offset the loss of enzyme activity through denaturation, so the system will show little or no overall loss of activity. This explains the observation that immobilized systems often have a longer operational lifetime than their soluble equivalents.

In addition, it is of interest that enzymes bound to natural cell membranes (phospholipid bilayers) within living cells will also probably demonstrate these effects, and immobilized systems thus provide useful models for the study of such membrane-bound proteins in living cells.

Immobilized enzymes at work

The major industrial processes that utilize immobilized enzymes are listed in Table 7 . Sales of immobilized enzymes peaked in 1990, when they accounted for about 20% of all industrial enzyme sales, almost entirely due to the use of glucose isomerase for the production of sweetening agents. Other commercial applications utilize penicillin acylase, fumarase, β–galactosidase and amino acid acylase. Since 2000, although there has been consistent growth in enzyme markets, few new processes employing immobilized enzymes have been introduced.

ProcessEnzymeProduction rate (ton year )
High-fructose corn syrup productionGlucose isomerase10
Acrylamide productionNitrile hydratase10
Transesterification of food oilsLipase10
Lactose hydrolysisLactase10
Semi-synthetic penicillin productionPenicillin acylase10
l-aspartic acid productionAspartase10
Aspartame productionThermolysin10

The following three examples highlight many of the biochemical, technological and economic considerations relating to the use of immobilized enzymes on a commercial and industrial scale.

Production of high-fructose syrup

Undoubtedly the most significant large-scale application of immobilized enzymes involves the production of high-fructose corn syrup (HFCS). Although most of the general public believe that sucrose is responsible for the ‘sweetness’ of food and drinks, there have been significant efforts to replace sucrose with alternative, and often cheaper, soluble caloric sweetening agents. HFCS is a soluble sweetener that has been used in many carbonated soft drinks since the 1980s, including brand-name colas such as Coca-Cola and Pepsi-Cola. HFCS is produced by the enzymatic digestion of starch derived from corn (maize). Developments in HFCS production have been most prominent in countries such as the U.S.A., which have a high capacity to produce starch in the form of corn, but which do not cultivate significant amounts of sugar cane or sugar beet, and must therefore import either the raw products (for processing) or the refined sugar (sucrose) itself.

Simple corn syrups can be manufactured by breaking down starch derived from corn using the enzyme glucoamylase alone or in combination with α-amylase. These enzymes are cheap and can be used in a soluble form. Since starch has to be extracted from corn at high temperatures (because starch has poor solubility at low temperatures and forms very viscous solutions), the process utilizes enzymes from thermophilic organisms, which have very high temperature optima. Simple corn syrup is therefore composed predominantly of glucose, which unfortunately has only 75% of the sweetness of sucrose. However, in order to make the syrup sweeter the enzyme glucose isomerase, which catalyses the following reaction, can be employed:

This enzyme (described previously in the section on properties and mechanisms of enzyme action) will produce a roughly 50:50 mixture of glucose and fructose at equilibrium, and since fructose has 150% of the sweetness of sucrose, this glucose:fructose mixture will have a similar level of sweetness to sucrose. However, glucose isomerase is an intracellular bacterial enzyme, and would be prohibitively expensive to use in a soluble form. This makes it an ideal candidate for use in an immobilized process.

The first glucose isomerase enzyme to be isolated was obtained from species of Pseudomona s in 1957, and more useful enzymes were isolated throughout the 1960s from species of Bacillus and Streptomyces. In 1967, the Clinton Corn Processing Company of Iowa, U.S.A. (later renamed CPC International) introduced a batch process that utilized an immobilized glucose isomerase enzyme, and by 1972 the company had developed a continuous process for the manufacture of HFCS containing 42% fructose using a glucose isomerase enzyme immobilized on a DEAE ion-exchange support.

During the late 1970s, advances in enzymology, process engineering and fractionation technology led to the production of syrups with a higher fructose content, and today HFCS containing 55% fructose is generally produced, and is commonly used in soft drinks, although 42% fructose syrups are still also produced for use in some processes, including the production of bakery foodstuffs.

In 2010, the U.S. production of HFCS was approximately 8 million metric tons, accounting for 37% of the U.S. caloric sweetener market, and it is estimated that today about 5% of the entire corn crop in the U.S.A. is used to produce HFCS.

Hydrolysis of lactose

Within the dairy industry the production of 1 kg of cheese requires about 10 litres of milk, and produces about 9 litres of whey as a waste product. Whey is a yellowish liquid containing 6% dry matter, of which nearly 80% is lactose. The enzyme lactase (β-galactosidase) may be used to break down lactose to its constituent monosaccharides, namely glucose and galactose, which are more soluble than lactose, and have potential uses as carbon sources in microbial fermentation, and can also be used as caloric sweeteners.

Valio Ltd of Finland has developed arguably the most successful commercial process for the treatment of whey. Using a lactase enzyme obtained from Aspergillus , immobilized by adsorption and cross-linked on to a support resin, whey syrups are produced that have been utilized as an ingredient in drinks, ice cream and confectionery products. The Aspergillus enzyme has an acid pH optimum of 3–5, and by operating at low pH the process avoids excessive microbial contamination. Treatment plants that utilize 600-litre columns have been built in Finland, and these are used to treat 80 000 litres of whey per day. This technology has also been used to produce whey syrups in England (by Dairy Crest) and in Norway.

Similar technology can also be used to remove lactose from milk. Lactose-free milk is produced for consumption by those who have lactose intolerance (a genetic condition), and also for consumption by pets such as cats, which are often unable to digest lactose easily. The first industrial processing facility to use immobilized lactase to treat milk was opened in 1975, when Centrale del Latte of Milan, Italy, utilized a batch process in which yeast ( Saccharomyces ) lactase, with a neutral pH optimum of 6–8, was immobilized within hollow permeable fibres. This process was capable of treating 10 000 litres of milk per day, and was operated at low temperature to prevent microbial contamination.

Production of semi-synthetic penicillins

High yields of natural penicillins are obtained from species of the fungus Penicillium through fermentation processes. However, over the years many microbial pathogens have become resistant to natural penicillins, and are now only treatable with semi-synthetic derivatives. These are produced through cleavage of natural penicillin, such that the G or V side chain is removed from the 6-aminopenicillanic acid (6-APA) nucleus of the molecule:

Thereafter, by attachment of a chemically different side chain, a semi-synthetic penicillin product (e.g. ampicillin, amoxicillin) can be formed. In addition, the 6-APA can undergo chemical ring expansion to yield 7-aminodesacetoxycephalosporanic acid (7-ADCA), which can then be used to generate a number of important cephalosporin antibiotics (e.g. cephalexin, cephradine, cefadroxil).

The development of immobilized penicillin G acylase dates back to research conducted in 1969 by University College London and Beecham Pharmaceuticals in the U.K. Penicillin G acylases are intracellular enzymes found in E. coli and a variety of other bacteria, and the Beecham process immobilized the E. coli enzyme on a DEAE ion-exchange support. Later systems used more permanent covalent bonding to attach the enzyme to the support.

In the 1980s and 1990s, world production of penicillins was dominated by European manufacturers, which accounted for production of around 30 000 tonnes of penicillin per annum, 75% of which was used for the manufacture of semi-synthetic penicillins and cephalosporins. However, over the past 10 years, due to increasing costs of labour, energy and raw materials, more bulk manufacturing has moved to the Far East, where China, Korea and India have become major producers. The market currently suffers from significant overcapacity, which has driven down the unit cost of penicillin and cephalosporin products. However, penicillins and cephalosporins still represent one of the world's major biotechnology markets, with annual sales of about £10 000 million, accounting for 65% of the entire global antibiotics market.

Enzymes in analysis

Enzymes have a wide variety of uses in analytical procedures. Their specificity and potency allow both detection and amplification of a target analyte. ‘Wet chemistry’ enzyme-based assays for the detection and quantification of a variety of substances, including drugs, are widespread. Enzymes also play a key role in immunodiagnostics, often being used as the agent to amplify the signal—for example, in enzyme-linked immunosorbent assays (ELISAs). Within DNA-fingerprinting technology, the enzyme DNA polymerase plays a key role in the amplification of DNA molecules in the polymerase chain reaction. However, ‘wet chemistry’ analytical methods are increasingly being replaced by the use of biosensors—that is, self-contained integrated devices which incorporate a biological recognition component (usually an immobilized enzyme) and an electrochemical detector (known as a transducer).

Much of the technological development of biosensors has been motivated by the need to measure blood glucose levels. In 2000, the World Health Organization estimated that over 170 million people had diabetes, and predicted that this figure will rise to over 360 million by 2030. In view of this, many companies have made significant investments in R&D programmes that have led to the availability of a wide variety of glucose biosensor devices.

In 1962, Leland Clark Jr coined the term ‘enzyme electrode’ to describe a device in which a traditional electrode could be modified to respond to other materials by the inclusion of a nearby enzyme layer. Clark's ideas became a commercial reality in 1975 with the successful launch of the Yellow Springs Instruments (YSI) model 23A glucose analyser. This device incorporated glucose oxidase together with a peroxide-sensitive electrode to measure the hydrogen peroxide (H 2 O 2 ) produced during the following reaction:

In this device, the rate of H 2 O 2 formation is a measure of the rate of the reaction, which depends on the concentration of glucose in solution, thus allowing the latter to be estimated.

As was discussed earlier, in enzyme-catalysed reactions the relationship between substrate concentration and reaction rate is not linear, but hyperbolic (as described by the Michaelis–Menten equation). This is also true for the glucose oxidase within a biosensor. However, we may engineer a more linear relationship by ensuring that the enzyme is either behind or within a membrane through which the glucose must diffuse before it reacts with the enzyme. This means that the system becomes diffusionally, rather than kinetically, limited, and the response is then more linearly related to the concentration of glucose in solution.

Over the years the YSI model 23A glucose analyser has been replaced by a range of much more advanced models. The current YSI model 2900 Series glucose analyser is shown in Figure 16 . This instrument has a 96-sample rack that enables batches of samples to be run, with the analysis of each sample taking less than a minute. The instrument can measure the glucose content of whole blood, plasma or serum, and requires only 10 μl of sample per analysis. The membrane-bound glucose oxidase typically only needs to be replaced every 3 weeks, thereby reducing the cost of analysis. These systems also offer advanced data-handling and data-storage facilities.

An external file that holds a picture, illustration, etc.
Object name is bse0590001i016.jpg

In addition, these instruments can be modified to analyse a wide variety of other substances of biological interest, simply by incorporating other oxidase enzymes into the membrane ( Table 8 ).

AnalyteEnzymeReaction
GlucoseGlucose oxidaseβ-D-glucose + O → gluconic acid + H O
AlcoholAlcohol oxidaseEthanol + O → acetaldehyde + H O
Lactic acidLactate oxidasel-lactate + O → pyruvate + H O
LactoseGalactose oxidaseLactose + O → galactose dialdehyde derivative + H O

To enable diabetic patients to take their own blood glucose measurements, small hand-held biosensors have also been developed, which are in fact technologically more advanced because the enzyme and transducer are more intimately linked on the sensor surface. The first device of this type was launched in 1986 by Medisense, and was based on technology developed in the U.K. at Cranfield and Oxford Universities. The ExacTech blood glucose meter was the size and shape of a pen, and used disposable electrode strips. This device was followed by a credit card-style meter in 1989. Such devices again rely on glucose oxidase as the biological component, but do not measure the reaction rate via the production (and detection) of H 2 O 2 . Instead they rely on direct measurement of the rate of electron flow from glucose to the electrode surface. The reactions that occur within this device may be summarized as follows:

and at the electrode surface:

where GO x -FAD represents the FAD redox centre of glucose oxidase in its oxidized form, and GO x -FADH 2 represents the reduced form.

Basically electrons are removed from the glucose molecules and passed via the enzyme to the ferrocene mediator, which then donates them to the working electrode surface, resulting in the generation of an electrical current that is directly proportional to the rate of oxidation of glucose, and thus proportional to the glucose concentration in the sample.

Medisense, whose only product was its blood glucose meter, was bought by Abbott Diagnostics in 1996, and Abbott-branded devices continued to use and develop this technology for some time.

In 1999, Therasense marketed a glucose meter that represented the next generation of sensing technology, and integrated the enzyme even more closely with the electrode. Originally developed by Adam Heller at the University of Texas in the 1990s, wired-enzyme electrodes do not rely on a soluble mediator such as the ferrocene used in the Medisense devices. Instead the enzyme is immobilized in an osmium-based polyvinyl imidazole hydrogel in which the electrons are passed from enzyme to electrode by a series of fixed electroactive osmium centres that shuttle the electrons onward in a process called ‘electron hopping.’

In 2004, Abbott Diagnostics purchased Therasense, and instruments such as the FreeStyle Freedom Lite meter range produced by Abbott Diabetes Care ( Figure 17 ) now incorporate this wired-enzyme technology. Devices of this type are highly amenable to miniaturization.

An external file that holds a picture, illustration, etc.
Object name is bse0590001i017.jpg

Continuous measuring devices are becoming increasingly available, and may well revolutionize the control of certain disease conditions. For example, with regard to diabetes, devices such as the FreeStyle Navigator range from Abbott Diabetes Care use the same wired-enzyme technology as that described earlier, but now incorporate this into a tiny filament about the diameter of a thin hypodermic needle. This is inserted approximately 5 mm under the skin to measure the glucose level in the interstitial fluid that flows between the cells. The unit is designed to remain in situ for up to 5 days, during which time it can measure the glucose concentration every minute. A wireless transmitter sends the glucose readings to a separate receiver anywhere within a 30-metre range, and this can then issue an early warning alarm to alert the user to a falling or rising glucose level in time for them to take appropriate action and avoid a hypoglycaemic or hyperglycaemic episode.

In addition, experimental units have already been developed that link continuous glucose biosensor measurement systems with pumps capable of gradually dispensing insulin such that the diabetic condition is automatically and reliably controlled, thereby avoiding the traditional peaks and troughs in glucose levels that occur with conventional glucose measurement and the intermittent administration of insulin.

Therefore, looking to the future, we may confidently expect to see the development of biosensor systems that can continuously monitor a range of physiologically important analytes and automatically dispense the required medication to alleviate the symptoms of a number of long-term chronic human illnesses.

Closing remarks

For the sake of conciseness, this guide has been limited to some of the basic principles of enzymology, together with an overview of the biotechnological applications of enzymes. It is important to understand the relationship between proteins and the nucleic acids (DNA and RNA) that provide the blueprint for the assembly of proteins within the cell. Genetic engineering is thus predominantly concerned with modifying the proteins that a cell contains, and genetic defects (in medicine) generally relate to the abnormalities that occur in the proteins within cells. Much of the molecular age of biochemistry is therefore very much focused on the study of the cell, its enzymes and other proteins, and their functions.

Abbreviations

This article is a reviewed, revised and updated version of the following ‘Biochemistry Across the School Curriculum’ (BASC) booklet: Teal A.R. & Wymer P.E.O., 1995: Enzymes and their Role in Technology. For further information and to provide feedback on this or any other Biochemical Society education resource, please contact gro.yrtsimehcoib@noitacude . For further information on other Biochemical Society publications, please visit www.biochemistry.org/publications .

Recommended reading and key publications

1. historically important landmark papers (in chronological order).

  • Takamine J. Process of making diastatic enzyme. 1894 U.S. Pat. 525,823. Describes the first commercial exploitation of semi-purified enzymes in the West. [ Google Scholar ]
  • Briggs G.E., Haldane J.B.S. A note on the kinetics of enzyme action. Biochem. J. 1925; 19 :338–339. A classic paper in which the steady-state assumption was introduced into the derivation of the Michaelis–Menten equation. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Koshland D.E., Jr Application of a theory of enzyme specificity to protein synthesis. Proc. Natl Acad. Sci. U.S.A. 1958; 44 :98–104. Describes the proposal of an ‘induced fit’ mechanism of substrate binding. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Clark L.C., Jr, Lyons C. Electrode systems for continuous monitoring in cardiovascular surgery. Ann. N.Y. Acad. Sci. 1962; 102 :29–45. Introduces the concept of a biosensor for measuring blood glucose levels during surgery. [ PubMed ] [ Google Scholar ]
  • Monod J., Wyman H., Changeux J.P. On the nature of allosteric transitions: a plausible model. J. Mol. Biol. 1965; 12 :88–118. Describes the ‘concerted’ model of transitions of allosteric proteins in which all constituent monomers are in either the T-state or the R-state. [ PubMed ] [ Google Scholar ]
  • Koshland D.E., Jr, Némethy G., Filmer D. Comparison of experimental binding data and theoretical models in proteins containing subunits. Biochemistry. 1966; 5 :365–385. Describes the ‘sequential’ model of transitions of allosteric proteins in which protein moves through hybrid structures with some monomers in the T-state and some in the R-state. [ PubMed ] [ Google Scholar ]
  • Updike S.J., Hicks G.P. The enzyme electrode. Nature. 1967; 214 :986–988. Describes the simplification of the electrochemical assay of glucose by immobilizing and thereby stabilizing the glucose oxidase enzyme. [ PubMed ] [ Google Scholar ]
  • Tramontano A., Janda K.D., Lerner R.A. Catalytic antibodies. Science. 1986; 234 :1566–1570. [ PubMed ] [ Google Scholar ]
  • Pollack S.J., Jacobs J.W., Schultz P.G. Selective chemical catalysis by an antibody. Science. 1986; 234 :1570–1573. The first reports of antibody proteins that demonstrate catalytic activity. [ PubMed ] [ Google Scholar ]
  • Johnson K.A., Goody R.S. The original Michaelis constant: translation of the 1913 Michaelis–Menten paper. Biochemistry. 2011; 50 :8264–8269. A modern translation, commentary and re-analysis of the original 1913 paper, Die Kinetik der Invertinwirkung. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Taylor A.I., Pinheiro V.B., Smola M.J., Morgunov A.S., Peak-Chew S., Cozens C., Weeks K.M., Herdewijn P., Holliger P. Catalysts from synthetic genetic polymers. Nature. 2015; 518 :427–430. Describes the first artificial enzymes to be created using synthetic biology. [ PMC free article ] [ PubMed ] [ Google Scholar ]

2. Enzyme principles

  • Changeux J.-P. 50 years of allosteric interactions: the twists and turns of the models. Nat. Rev. Mol. Cell Biol. 2013; 14 :819–829. [ PubMed ] [ Google Scholar ]
  • Kamata K., Mitsuya M., Nishimura T., Eiki J., Nagata Y. Structural basis for allosteric regulation of the monomeric allosteric enzyme human glucokinase. Structure. 2004; 12 :429–438. [ PubMed ] [ Google Scholar ]

3. Enzyme applications

  • Adrio J.L., Demain A.L. Microbial enzymes: tools for biotechnological processes. Biomolecules. 2014; 4 :117–139. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Clarke S.F., Foster J.R. A history of blood glucose meters and their role in self-monitoring of diabetes mellitus. Br. J. Biomed. Sci. 2012; 69 :83–93. [ PubMed ] [ Google Scholar ]
  • Fernandes P. Enzymes in food processing: a condensed overview on strategies for better biocatalysts. Enzyme Res. 2010; 2010 :862537. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Vashist S.K., Zheng D., Al-Rubeaan K., Luong J.H.T., Sheu F.-S. Technology behind commercial devices for blood glucose monitoring in diabetes management: a review. Anal. Chim. Acta. 2011; 703 :124–136. [ PubMed ] [ Google Scholar ]
  • Vellard M. The enzyme as drug: application of enzymes as pharmaceuticals. Curr. Opin. Biotechnol. 2003; 14 :444–450. [ PubMed ] [ Google Scholar ]
  • Woodley J.M. New opportunities for biocatalysis: making pharmaceutical processes greener. Trends Biotechnol. 2008; 26 :321–327. [ PubMed ] [ Google Scholar ]

4. Useful textbooks

  • Bisswanger H. Enzyme Kinetics: Principles and Methods. 2nd edn. Weinheim, Germany: Wiley-VCH; 2008. Available online and as hard copy. A user-friendly and comprehensive treatise on enzyme kinetics. [ Google Scholar ]
  • Buchholz K., Kasche V., Bornscheuer U.T. Biocatalysts and Enzyme Technology. 2nd edn. Weinheim, Germany: Wiley-VCH; 2012. Best-selling textbook that provides an instructive and comprehensive overview of our current knowledge of biocatalysis and enzyme technology. [ Google Scholar ]
  • Copeland R.A. In: Evaluation of Enzyme Inhibitors in Drug Discovery: A Guide for Medicinal Chemists and Pharmacologists. 2nd edn. Hoboken NJ., editor. John Wiley & Sons, Inc; 2013. Provides thorough coverage of both the principles and applications of enzyme inhibitors. [ PubMed ] [ Google Scholar ]
  • McGrath M.J., Scanaill C.N. Sensor Technologies: Healthcare, Wellness and Environmental Applications. New York: Apress Media, LLC; 2014. Available online. Covers sensor technologies and their clinical applications, together with broader applications that are relevant to wellness, fitness, lifestyle and the environment. [ Google Scholar ]
  • Trevan M.D. Immobilized Enzymes: An Introduction and Applications in Biotechnology. Chichester: John Wiley & Sons; 1980. An older text, and difficult to find except in libraries, but it provides an introductory text for non-experts, and as yet there is no other book that fulfils this role. [ Google Scholar ]
  • Whitehurst R.J., van Oort M. Enzymes in Food Technology. 2nd edn. Chichester: Wiley-Blackwell; 2009. Provides comprehensive coverage of the widespread use of enzymes in food-processing improvement and innovation. [ Google Scholar ]

FREE K-12 standards-aligned STEM

curriculum for educators everywhere!

Find more at TeachEngineering.org .

  • TeachEngineering
  • The Power of Enzymes!

Lesson The Power of Enzymes!

Grade Level: 5 (4-6)

Time Required: 45 minutes

Lesson Dependency: None

Subject Areas: Biology, Chemistry, Life Science, Measurement, Physical Science

  • Print lesson and its associated curriculum

Activities Associated with this Lesson Units serve as guides to a particular content or subject area. Nested under units are lessons (in purple) and hands-on activities (in blue). Note that not all lessons and activities will exist under a unit, and instead may exist as "standalone" curriculum.

  • Clean, Green Washing Machine Challenge
Lesson Activity

TE Newsletter

Engineering connection, learning objectives, worksheets and attachments, more curriculum like this, introduction/motivation, associated activities, lesson closure, vocabulary/definitions, user comments & tips.

Engineers team up to tackle global challenges

Laundry is a real-world problem to which we can all relate. Engineers work to solve all types of problems, including how to create chemicals that help us in our daily lives. In the case of cleaning agents used in household products such as laundry detergent, chemical engineers use their knowledge of chemistry to engineer formulas. Engineers look to a variety of sources to solve problems and the human body is an excellent source of inspiration. For example, enzymes are present in the digestive system of the human body and these digestive enzymes—amylase, protease and lipase—have a specific function of breaking down food. Engineers use this scientific knowledge to create a laundry detergent using these enzymes which break down and remove a variety of different laundry stains created by foods.

After this lesson, students should be able to:

  • Explain the role of enzymes in digestion.
  • Explain why enzymes are used in laundry detergent.

Educational Standards Each TeachEngineering lesson or activity is correlated to one or more K-12 science, technology, engineering or math (STEM) educational standards. All 100,000+ K-12 STEM standards covered in TeachEngineering are collected, maintained and packaged by the Achievement Standards Network (ASN) , a project of D2L (www.achievementstandards.org). In the ASN, standards are hierarchically structured: first by source; e.g. , by state; within source by type; e.g. , science or mathematics; within type by subtype, then by grade, etc .

Ngss: next generation science standards - science.

View aligned curriculum

Do you agree with this alignment? Thanks for your feedback!

International Technology and Engineering Educators Association - Technology

State standards, florida - english.

Have you ever wondered how food gets digested in your body? Our body needs food for energy and to help our cells work.  When we eat food, it travels through our digestive system. Along the way, the food gets broken down by special helpers called enzymes.  Enzymes are proteins that help the body digest food. There are many kinds of enzymes in our body, but there are three main enzymes that help us break down most of the food we eat. The first one, called amylase , helps us break down carbohydrates like pasta or cereal. The second one, protease , helps us break down proteins like meat or cheese. The third enzyme lipase helps us break down fats like oils or butter. 

Now we are going to read a book to help us better understand how the digestive system works. While we read this book, I want you to notice whenever the word 'enzymes' comes up in the book.

(Read the book: Amazing Body Systems: Digestive System , by Karen Latchana Kenney or A Journey Through the Digestive System with Max Axiom, Super Scientist , by Emily Sohn.)

Now let’s go back to the pages where enzymes were mentioned.   Where were the enzymes in the body? How do they work in digestion?  (Repeat for each occurrence.)

When you eat food, do you ever spill and make a stain on your clothes? (Prompt students to say yes or no to this question.)  What do you do when you get a stain on your clothes? (Possible answers: someone in my family helps me wash it out, or I wash it out, etc.) . Some laundry detergents include enzymes in their active ingredients. 

Based on what you now know from the reading and our discussion, why do you think laundry detergents would include enzymes in their ingredients? (Possible answers: enzymes work on stains, enzymes digest food stains, etc.) .

Lesson Background and Concepts for Teachers

Teachers need to know some background information on enzymes role in digestion and how enzymes are used in laundry detergents. Please use the books and links below for background information:

  • A Journey Through the Digestive System with Max Axiom, Super Scientist, by Emily Sohn
  • Amazing Body Systems: Digestive System, by Karen Latchana Kenney
  • Video: What are the enzymes of the digestive system? https://www.youtube.com/watch?v=Ej-WkJEaXNU
  • Video: Biological detergents, https://www.youtube.com/watch?v=23n4RNwptDg
  • Video: Enzymes, https://www.youtube.com/watch?v=qgVFkRn8f10
  • What are digestive enzymes? https://www.wisegeek.com/what-are-digestive-enzymes.htm

Watch this activity on YouTube

We can learn a lot from the human body. Today we learned that we can break down stains with the same enzymes that our bodies use to break down foods.

amylase: An enzyme that breaks down carbohydrates.

detergent: A substance that cleans.

digestion: The process of breaking down food by mechanical and enzymatic action to be used by the body.

enzyme: A protein that speeds up a chemical reaction.

lipase: An enzyme that breaks down fats (lipids).

protease: An enzyme that breaks down proteins.

Pre-Lesson Assessment:

What we - Enzymes What we - Enzymes What we - Enzymes
 

Lesson Summary Assessment:

What we - Enzymes What we - Enzymes What we - Enzymes
is a protein that speeds up a chemical reaction. work in our digestive system is an enzyme that breaks down carbohydrates is an enzyme that breaks down proteins is an enzyme that breaks down fats (lipids).

Exit Ticket: Have students complete the Exit Ticket .

Lesson Extension Activities

Using the Laundry Test Lab Worksheet , students can test 2 or 3 different enzymatic laundry detergents on some typical foods to test different stains to see which is the hardest to get out or easiest.  Students can have fun squeezing the bottles on to clean rags to make the stains and then testing detergents to see which one is most effective at getting out the stains.  Students could use beakers or graduated cylinders to measure the water and detergent and pipettes to transfer the water and detergent onto the stains. Students can use rulers to measure the size of the stain before and after using the detergent and water.

Students performing the lesson extension activity using detergent and stained clothing.

This lesson introduces students to the main parts of the digestive system and how they interact. In addition, students learn about some of the challenges astronauts face when eating in outer space. Engineers figure out how to deal with such challenges.

preview of 'Digestive System' Lesson

To reinforce students' understanding of the human digestion process, the functions of several stomach and small intestine fluids are analyzed, and the concept of simulation is introduced through a short, introductory demonstration of how these fluids work. Students learn what simulation means and ho...

preview of 'Digestion Simulation' Lesson

Contributors

Supporting program, acknowledgements.

This curriculum was based upon work supported by the National Science Foundation under RET grant no. EEC 1711543— Engineering for Biology: Multidisciplinary Research Experiences for Teachers in Elementary Grades (MRET) through the College of Engineering at the University of Florida. Any opinions, findings, and conclusions or recommendations expressed in this material are those of the authors and do not necessarily reflect the views of the National Science Foundation.

Last modified: October 26, 2020

Practical: Investigating Temperature & Enzyme Activity ( Edexcel IGCSE Biology )

Revision note.

Lára

Biology Lead

Practical: Enzymes & Temperature

  • Amylase is an enzyme that digests starch (a polysaccharide of glucose) into maltose (a disaccharide of glucose)
  • Spotting tile
  • Measuring cylinder
  • Thermometer
  • Starch solution
  • Amylase solution
  • Add 5cm 3 starch solution to a test tube and heat to a set temperature using beaker of water with a Bunsen burner
  • Add a drop of Iodine to each of the wells of a spotting tile
  • Use a syringe to add 2cm 3   a mylase to the starch solution and mix well
  • Every minute, transfer a droplet of solution to a new well of iodine solution (which should turn blue-black)
  • Repeat this transfer process until the iodine solution stops turning blue-black (this means the amylase has broken down all the starch )
  • Record the time taken for the reaction to be completed
  • Repeat the investigation for a range of temperatures (from 20°C to 60°C)

Investigating the effect of temperature on enzyme activity, IGCSE & GCSE Biology revision notes

Investigating the effect of temperature on enzyme activity

Results and Analysis

  • Amylase is an enzyme which breaks down starch
  • The quicker the reaction is completed, the faster the enzyme is working
  • This is because the enzyme is working at its fastest rate and has digested all the starch
  • This is because the amylase enzyme is working slowly due to low kinetic energy and few collisions between the amylase and the starch
  • This is because the amylase enzyme has become denatured and so can no longer bind with the starch or break it down

Limitations

  • Note that there are several different ways in which the temperature could be controlled. The method described above is not very precise, an improvement would be to use water baths kept at each temperature
  • The starch and amylase solutions that need to be used should be placed in a water bath and allowed to reach the temperature (using a thermometer to check) before being used
  • A solution containing starch will be darker than a solution containing glucose (as a result of the colour change of iodine)
  • The absorbance or transmission of light through the coloured solution can be measured using a colorimeter

Applying CORMS to practical work

  • When working with practical investigations, remember to consider your CORMS evaluation

CORMS evaluation, downloadable AS & A Level Biology revision notes

CORMS evaluation

  • C - We are changing the temperature in each repeat
  • O - This is not relevant to this investigation as we aren't using an organism
  • R - We will repeat the investigation several times to make sure our results are reliable
  • M1 - We will measure the time taken
  • M2 - for the iodine to stop turning black
  • S - We will control the concentration and volume of starch solution, iodine and amylase used in the investigation

Describing and explaining experimental results for enzyme experiments is a common type of exam question so make sure you understand what is happening and can relate this to changes in the active site of the enzyme when it has denatured, or if it is a low temperature , relate it to the amount of kinetic energy the molecules have.

You've read 0 of your 0 free revision notes

Get unlimited access.

to absolutely everything:

  • Downloadable PDFs
  • Unlimited Revision Notes
  • Topic Questions
  • Past Papers
  • Model Answers
  • Videos (Maths and Science)

Join the 100,000 + Students that ❤️ Save My Exams

the (exam) results speak for themselves:

Did this page help you?

Author: Lára

Lára graduated from Oxford University in Biological Sciences and has now been a science tutor working in the UK for several years. Lára has a particular interest in the area of infectious disease and epidemiology, and enjoys creating original educational materials that develop confidence and facilitate learning.

experiment about enzymes

Faraday Discussions

Enzyme-modified pt nanoelectrodes for glutamate detection.

We present here a glutamate oxidase (GluOx)-modified platinum (Pt) nanoelectrode with a planar geometry for glutamate detection. The Pt nanoelectrode was characterized using electrochemistry and scanning electron microscopy (SEM). The radius of the Pt nanoelectrode measured using SEM is ~210 nm. GluOx-modified Pt nanoelectrodes were generated by dip coating GluOx on the Pt nanoelectrode in a solution of 0.9% (wt%) bovine serum albumin (BSA), 0.126% (wt%) glutaraldehyde, and 100 U/mL GluOx. An increase in current was observed at +0.7 V vs. Ag/AgCl/1M KCl with adding increasing concentrations of glutamate. A two-sample t-test results showed that there is a significant difference for current at +0.7 V between the blank and the added lowest glutamate concentration, as well as between adjacent glutamate concentrations, confirming that the increase in current is related to the increased glutamate concentration. The experimental current-concentration curve of glutamate detection fitted well to the theoretical Michaelis-Menten curve. At the low concentration range (50 μM to 200 μM), a linear relationship between the current and glutamate concentration was observed. The Michaelis-Menten constants of Imax and Km were calculated to be 1.093 pA and 0.227 mM, respectively. Biosensor efficiency (the ratio of glutamate sensitivity to H2O2 sensitivity) is calculated to be 57.9%. Enzact (Imax /H2O2 sensitivity, an indicator of the amount of enzyme loaded on the electrode) of the GluOx-modified Pt nanoelectrode is 0.243 mM. We further compared the sensitivity of a GluOx-modified Pt nanoelectrode with a GluOx-modified carbon fiber microelectrode (7-μm diameter and a sensing length of ~350 μm). Glutamate detection on the GluOx-modified carbon fiber microelectrode fitted well to a Michaelis-Menten like response. Based on the fitting, the GluOx-modified carbon fiber microelectrode exhibited an Imax of 0.689 nA and a Km of 301.2 μM towards glutamate detection. The best linear range of glutamate detection on the GluOx-modified carbon fiber microelectrode is from 50 μM to 150 μM Glutamate. GluOx-modified carbon fiber microelectrode exhibited a higher potential requirement for glutamate detection comparing to the GluOx-modified Pt nanoelectrode.

  • This article is part of the themed collection: New horizons in nanoelectrochemistry

Supplementary files

  • Supplementary information PDF (632K)

Article information

Download citation, permissions.

experiment about enzymes

P. Xu, H. D. Jetmore, R. Chen and M. Shen, Faraday Discuss. , 2024, Accepted Manuscript , DOI: 10.1039/D4FD00138A

To request permission to reproduce material from this article, please go to the Copyright Clearance Center request page .

If you are an author contributing to an RSC publication, you do not need to request permission provided correct acknowledgement is given.

If you are the author of this article, you do not need to request permission to reproduce figures and diagrams provided correct acknowledgement is given. If you want to reproduce the whole article in a third-party publication (excluding your thesis/dissertation for which permission is not required) please go to the Copyright Clearance Center request page .

Read more about how to correctly acknowledge RSC content .

Social activity

Search articles by author.

This article has not yet been cited.

Advertisements

IMAGES

  1. Enzyme Activity

    experiment about enzymes

  2. B2 D) Enzyme Investigations

    experiment about enzymes

  3. Potato Catalase Enzyme Experiment (UPDATE)

    experiment about enzymes

  4. Experiment in Enzymes || Biochemistry

    experiment about enzymes

  5. Enzyme Experiments: Factors That Affect Enzyme Activity

    experiment about enzymes

  6. Lab 14 Enzyme activity experiment

    experiment about enzymes

VIDEO

  1. Isotonix digestive enzymes experiment

  2. Immobilised Enzymes.wmv

  3. ENZYMES

  4. 🤯DIGESTIVE ENZYMES EXPERIMENT 🤯 #ytshorts #science #lifesciencefacts #humanbiology #scienceandfun

  5. Enzymes for Cutting DNA Restriction Endonucleases

  6. digestive enzymes: Allerzyme Vs Essentialzymes-4

COMMENTS

  1. Exploring Enzymes

    The lower the activation energy of a reaction, the faster it takes place. If the activation energy is too high, the reaction does not occur. Enzymes have the ability to lower the activation energy of a chemical reaction by interacting with its reactants. Each enzyme has an active site, which is where the reaction takes place (Figure 1).

  2. Explore Enzyme Activity with Toothpicks

    An enzyme's activity tells you how well an enzyme performs its function and how fast the reaction takes place. The study of how enzymes change the rate at which a chemical reaction occurs is called enzyme kinetics. Enzymatic activity, or the reaction rate of the enzymatic reaction, is usually measured by doing an enzyme assay. ...

  3. Experiments on Enzyme Activity

    The below mentioned article includes a collection of seven experiments on enzyme activity. 1. Experiment to demonstrate the activity of enzymes: Requirements: Benzidine solution, razor, thin sections of actively growing root (or germinating seeds or germinating pollen grains), phosphate buffer, hydrogen peroxide (1%), ammonium chloride (5% ...

  4. Exploring Enzymes

    In this science activity you will investigate one of these enzymes, called catalase, to find out how it helps to protect your body from damage. Background Enzymes are essential for our survival ...

  5. PDF Experiment 10

    Experiment 10 - Enzymes. Enzymes are proteins that act as catalysts for biological reactions. Enzymes, like all catalysts, speed up reactions without being used up themselves. They do this by lowering the activation energy of a reaction. All biochemical reactions are catalyzed by enzymes. Since enzymes are proteins, they can be denatured in a ...

  6. Enzymes

    Enzymes are substrate specific, meaning that they catalyze only specific reactions. For example, proteases (enzymes that break peptide bonds in proteins) will not work on starch (which is broken down by the enzyme amylase). Notice that both of these enzymes end in the suffix -ase. This suffix indicates that a molecule is an enzyme.

  7. Testing for catalase enzymes

    The enzyme catalase can speed up (catalyse) this reaction. In this practical, students investigate the presence of enzymes in liver, potato and celery by detecting the oxygen gas produced when hydrogen peroxide decomposes. The experiment should take no more than 20-30 minutes.

  8. Enzyme structure and function (article)

    The answer is: enzymes! Enzymes are life's great facilitators. They create the conditions needed for biochemical reactions to happen fast. The general name that chemists use for a chemical entity that increases the speed of a reaction is a "catalyst." ... This magnet thought experiment is a good approximation of what happens with real ...

  9. PDF Simple Enzyme Experiments

    level at which they are used. The following enzymes are included: amylase, catalase, catecholase, invertase, papain, pectinase, pepsin, and rennin. Except for the pepsin experiment, all experiments can be completed during a 2- to 3-hour laboratory period. The time for each individual experiment varies from "instant" results with catalase

  10. Everything About Enzymes!! (and a free lab!)

    Enzymes are biological catalysts that speed up the chemical reactions of the cell. Enzymes are proteins. Enzymatic reactions occur faster and at lower temperatures because enzymes lower the activation energy for that chemical reaction. Enzymes are never consumed or used up during the reaction. They can do their job over and over again.

  11. Enzymes review (article)

    Enzymes are reusable. Enzymes are not reactants and are not used up during the reaction. Once an enzyme binds to a substrate and catalyzes the reaction, the enzyme is released, unchanged, and can be used for another reaction. This means that for each reaction, there does not need to be a 1:1 ratio between enzyme and substrate molecules.

  12. Enzyme Experiments: Factors That Affect Enzyme Activity

    This video shows how to set up two experiments - one looking at how pH affects the activity of catalase enzyme, and the other looking at how temperature affe...

  13. Enzyme Lab

    In this lab, you will study an enzyme that is found in the cells of many living tissues. The name of the enzyme is catalase; it speeds up a reaction which breaks down hydrogen peroxide, a toxic chemical, into 2 harmless substances--water and oxygen. Light can also break down H 2 O 2 which is why the chemical is sold in dark containers.

  14. Enzymes (video)

    Transcript. Enzymes are biological catalysts that speed up reactions. The active site is where substrates bind to the enzyme. Induced fit occurs when the enzyme changes shape to better accommodate substrates, facilitating the reaction. Enzymes can be used multiple times and are affected by factors such as temperature and pH. Created by Sal Khan.

  15. 8 Enzymes

    8 Enzymes. 8. Enzymes. Enzymes are macromolecular biological catalysts. The molecules upon which enzymes may act are called substrates and the enzyme converts the substrates into different molecules known as products. Almost all metabolic processes in the cell need enzyme catalysis in order to occur at rates fast enough to sustain life.

  16. Enzymes: principles and biotechnological applications

    Enzymes are potent catalysts. The enormous catalytic activity of enzymes can perhaps best be expressed by a constant, k cat, that is variously referred to as the turnover rate, turnover frequency or turnover number.This constant represents the number of substrate molecules that can be converted to product by a single enzyme molecule per unit time (usually per minute or per second).

  17. Enzyme Experiment

    This week, we are talking enzymes. In this enzyme experiment, you will get to see enzymes in action and an experiment challenge for you to do on your own. I hold a master's degree in child development and early education and am working on a post-baccalaureate in biology. I spent 15 years working for a biotechnology company developing IT ...

  18. Enzymes and the active site (article)

    Related to this is enzyme regulation where modification of an amino acid remote from the active site can control the activity of the enzyme. Furthermore, in some enzymes there is a second binding site and when something, such as an inhibitor, binds to that site it changes the shape of the active site, also controlling the enzyme.

  19. The Power of Enzymes!

    Enzymes are proteins that help the body digest food. There are many kinds of enzymes in our body, but there are three main enzymes that help us break down most of the food we eat. The first one, called amylase, helps us break down carbohydrates like pasta or cereal. The second one, protease, helps us break down proteins like meat or cheese.

  20. Lab 14 Enzyme activity experiment

    The decomposition of hydrogen peroxide with the enzyme catalase (in a banana) is studied. The reaction is done with various pHs, temperatures, and substrate...

  21. Practical: Investigating Temperature & Enzyme Activity

    Describing and explaining experimental results for enzyme experiments is a common type of exam question so make sure you understand what is happening and can relate this to changes in the active site of the enzyme when it has denatured, or if it is a low temperature, relate it to the amount of kinetic energy the molecules have.

  22. Enzyme Activity Lab Report

    Continuing, in the tests conducted, the higher the temperature the faster the reactions happened. Next, its time to move on to the societal benefits for this experiment. Investigating the effect of enzymes based on temperature could help us further understand body processes, because enzymes are a big part of body functions and reactions.

  23. Enzyme-modified Pt nanoelectrodes for glutamate detection

    Enzact (Imax /H2O2 sensitivity, an indicator of the amount of enzyme loaded on the electrode) of the GluOx-modified Pt nanoelectrode is 0.243 mM. We further compared the sensitivity of a GluOx-modified Pt nanoelectrode with a GluOx-modified carbon fiber microelectrode (7-μm diameter and a sensing length of ~350 μm).