Practical Biology

A collection of experiments that demonstrate biological concepts and processes.

redox reaction and catalase activity experiment

Observing earthworm locomotion

redox reaction and catalase activity experiment

Practical Work for Learning

redox reaction and catalase activity experiment

Published experiments

Investigating an enzyme-controlled reaction: catalase and hydrogen peroxide concentration, class practical or demonstration.

Hydrogen peroxide ( H 2 O 2 ) is a by-product of respiration and is made in all living cells. Hydrogen peroxide is harmful and must be removed as soon as it is produced in the cell. Cells make the enzyme catalase to remove hydrogen peroxide.

This investigation looks at the rate of oxygen production by the catalase in pureed potato as the concentration of hydrogen peroxide varies. The oxygen produced in 30 seconds is collected over water. Then the rate of reaction is calculated.

Lesson organisation

You could run this investigation as a demonstration at two different concentrations, or with groups of students each working with a different concentration of hydrogen peroxide. Individual students may then have time to gather repeat data. Groups of three could work to collect results for 5 different concentrations and rotate the roles of apparatus manipulator, result reader and scribe. Collating and comparing class results allows students to look for anomalous and inconsistent data.

Apparatus and Chemicals

For each group of students:.

Pneumatic trough/ plastic bowl/ access to suitable sink of water

Conical flask, 100 cm 3 , 2

Syringe (2 cm 3 ) to fit the second hole of the rubber bung, 1

Measuring cylinder, 100 cm 3 , 1

Measuring cylinder, 50 cm 3 , 1

Clamp stand, boss and clamp, 2

Stopclock/ stopwatch

For the class – set up by technician/ teacher:

Hydrogen peroxide, range of concentrations, 10 vol, 15 vol, 20 vol, 25 vol, and 30 vol, 2 cm 3 per group of each concentration ( Note 1 )

Pureed potato, fresh, in beaker with syringe to measure at least 20 cm 3 , 20 cm 3 per group per concentration of peroxide investigated ( Note 2 )

Rubber bung, 2-holed, to fit 100 cm 3 conical flasks – delivery tube in one hole (connected to 50 cm rubber tubing)

Health & Safety and Technical notes

Wear eye protection and cover clothing when handling hydrogen peroxide. Wash splashes of pureed potato or peroxide off the skin immediately. Be aware of pressure building up if reaction vessels become blocked. Take care inserting the bung in the conical flask – it needs to be a tight fit, so push and twist the bung in with care.

Read our standard health & safety guidance

1 Hydrogen peroxide: (See CLEAPSS Hazcard) Solutions less than 18 vol are LOW HAZARD. Solutions at concentrations of 18-28 vol are IRRITANT. Take care when removing the cap of the reagent bottle, as gas pressure may have built up inside. Dilute immediately before use and put in a clean brown bottle, because dilution also dilutes the decomposition inhibitor. Keep in brown bottles because hydrogen peroxide degrades faster in the light. Discard all unused solution. Do not return solution to stock bottles, because contaminants may cause decomposition and the stock bottle may explode after a time.

2 Pureed potato may irritate some people’s skin. Make fresh for each lesson, because catalase activity reduces noticeably over 2/3 hours. You might need to add water to make it less viscous and easier to use. Discs of potato react too slowly.

3 If the bubbles from the rubber tubing are too big, insert a glass pipette or glass tubing into the end of the rubber tube.

SAFETY: Wear eye protection and protect clothing from hydrogen peroxide. Rinse splashes of peroxide and pureed potato off the skin as quickly as possible.

Preparation

a Make just enough diluted hydrogen peroxide just before the lesson. Set out in brown bottles ( Note 1 ).

b Make pureed potato fresh for each lesson ( Note 2 ).

c Make up 2-holed bungs as described in apparatus list and in diagram.

Apparatus for investigation of an enzyme-controlled reaction

Investigation

d Use the large syringe to measure 20 cm 3 pureed potato into the conical flask.

e Put the bung securely in the flask – twist and push carefully.

f Half-fill the trough, bowl or sink with water.

g Fill the 50 cm 3 measuring cylinder with water. Invert it over the trough of water, with the open end under the surface of the water in the bowl, and with the end of the rubber tubing in the measuring cylinder. Clamp in place.

h Measure 2 cm 3 of hydrogen peroxide into the 2 cm 3 syringe. Put the syringe in place in the bung of the flask, but do not push the plunger straight away.

i Check the rubber tube is safely in the measuring cylinder. Push the plunger on the syringe and immediately start the stopclock.

j After 30 seconds, note the volume of oxygen in the measuring cylinder in a suitable table of results. ( Note 3 .)

k Empty and rinse the conical flask. Measure another 20 cm 3 pureed potato into it. Reassemble the apparatus, refill the measuring cylinder, and repeat from g to j with another concentration of hydrogen peroxide. Use a 100 cm 3 measuring cylinder for concentrations of hydrogen peroxide over 20 vol.

l Calculate the rate of oxygen production in cm 3 /s.

m Plot a graph of rate of oxygen production against concentration of hydrogen peroxide.

Shop Experiment Enzyme Action: Testing Catalase Activity Experiments​

Enzyme action: testing catalase activity.

Experiment #2B from Advanced Biology with Vernier

redox reaction and catalase activity experiment

Introduction

Many organisms can decompose hydrogen peroxide (H 2 O 2 ) enzymatically. Enzymes are globular proteins, responsible for most of the chemical activities of living organisms. They act as catalysts, substances that speed up chemical reactions without being destroyed or altered during the process. Enzymes are extremely efficient and may be used over and over again. One enzyme may catalyze thousands of reactions every second. Both the temperature and the pH at which enzymes function are extremely important. Most organisms have a preferred temperature range in which they survive, and their enzymes most likely function best within that temperature range. If the environment of the enzyme is too acidic or too basic, the enzyme may irreversibly denature , or unravel, until it no longer has the shape necessary for proper functioning.

H 2 O 2 is toxic to most living organisms. Many organisms are capable of enzymatically destroying the H 2 O 2 before it can do much damage. H 2 O 2 can be converted to oxygen and water, as follows:

2{\text{ }}{{\text{H}}_{\text{2}}}{{\text{O}}_{\text{2}}} \to {\text{2 }}{{\text{H}}_{\text{2}}}{\text{O + }}{{\text{O}}_{\text{2}}}

Although this reaction occurs spontaneously, enzymes increase the rate considerably. At least two different enzymes are known to catalyze this reaction: catalase, found in animals and protists, and peroxidase , found in plants. A great deal can be learned about enzymes by studying the rates of enzyme-catalyzed reactions.

In this experiment, you will

  • Use a Gas Pressure Sensor to measure the production of oxygen gas as hydrogen peroxide is destroyed by the enzyme catalase or peroxidase at various enzyme concentrations.
  • Measure and compare the initial rates of reaction for this enzyme when different concentrations of enzyme react with H 2 O 2 .
  • Measure the production of oxygen gas as hydrogen peroxide is destroyed by the enzyme catalase or peroxidase at various temperatures.
  • Measure and compare the initial rates of reaction for the enzyme at each temperature.
  • Measure the production of oxygen gas as hydrogen peroxide is destroyed by the enzyme catalase or peroxidase at various pH values.
  • Measure and compare the initial rates of reaction for the enzyme at each pH value.

Sensors and Equipment

This experiment features the following sensors and equipment. Additional equipment may be required.

redox reaction and catalase activity experiment

Ready to Experiment?

Ask an expert.

Get answers to your questions about how to teach this experiment with our support team.

Purchase the Lab Book

This experiment is #2B of Advanced Biology with Vernier . The experiment in the book includes student instructions as well as instructor information for set up, helpful hints, and sample graphs and data.

redox reaction and catalase activity experiment

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

redox reaction and catalase activity experiment

  • Back to parent navigation item
  • Collections
  • Sustainability in chemistry
  • Simple rules
  • Teacher well-being hub
  • Women in chemistry
  • Global science
  • Escape room activities
  • Decolonising chemistry teaching
  • Teaching science skills
  • Get the print issue
  • RSC Education

Three cartoons: a female student thinking about concentration, a male student in a wheelchair reading Frankenstein and a female student wearing a headscarf and safety goggles heating a test tube on a bunsen burner. All are wearing school uniform.

  • More navigation items

Investigating activation energies

By Derek Cheung 2007-01-01T00:00:00+00:00

A challenge for post-16 students to investigate the activation energies of the enzyme-catalysed and the inorganic-catalysed decomposition of hydrogen peroxide

  • Students extract catalase from fresh lettuce leaves to use as a catalyst for the decomposition of H 2 O 2
  • Reaction rates are determined by measuring the O 2 evolved over water in an inverted cylinder

Post-16 students are challenged to design and carry out an investigation to compare the activation energies of the enzyme-catalysed and the inorganic-catalysed decomposition of hydrogen peroxide.

In our discussions with secondary school teachers in Hong Kong we are told that students, aged 16-17, generally find it difficult to plan their own experiments to determine activation energies for reactions. To address this problem we have devised a laboratory-based investigation which challenges students to compare the activation energies for the catalase-catalysed and iodide-catalysed decomposition of hydrogen peroxide. This investigation should also help them to understand the different effects of biological and inorganic catalysts on activation energy. To make the investigation more authentic, simulating the work of a biochemist for example, students extract catalase from a natural source, in this case lettuce leaves. Our trials demonstrated that lettuce produces a suitable concentration of catalase for this kind of investigation. 

The challenge

The students are given a short introduction to the reaction under investigation, and their task. 

Kinetic studies play a significant role in elucidating the mechanisms of a reaction. A catalyst offers an alternative reaction pathway with a lower activation energy than the uncatalysed pathway, thus increasing the proportion of molecules that have enough energy to form products. The decomposition of hydrogen peroxide into water and oxygen can be catalysed by inorganic substances such as iodide ion, iron(II) ion, and manganese(IV) oxide. 

2H 2 O 2  → 2H 2 O + O 2

This reaction also takes place in human cells. Metabolism of fats and carbohydrates produces toxic hydrogen peroxide as a byproduct. One of the harmful effects of hydrogen peroxide is that it can damage DNA and cell membranes, resulting in ageing. Fortunately, the enzyme catalase, a biological catalyst present in our cells, speeds up the decomposition of hydrogen peroxide into harmless water and oxygen molecules. Catalase is also present in some vegetables and fruits. Examples are onion, carrot, celery, lettuce and papaya. 

Image - feature - cheung - main2

Extracting the enzyme catalase from lettuce

Imagine that you work as a biochemist and want to investigate the catalase activity in lettuce. You have to design and carry out an investigation to compare the activation energies for the catalase-catalysed and iodide-catalysed decomposition of hydrogen peroxide. Based on the activation energies, you can determine whether iodide or catalase is the more efficient catalyst. 

Students are expected to get approval of their experimental procedure from their teacher, who must also discuss any safety precautions with them before they begin. 

Enzymologists measure reaction rate by following the catalase-catalysed decomposition of hydrogen peroxide by using modern analytical methods such as UV spectrophotometry and chemiluminescence. 1  Older methods include titration with potassium manganate(VII) (KMnO 4 ) and colorimetry. Owing to constraints on equipment and time, our students determined the reaction rate by collecting and measuring the oxygen evolved over water in an inverted cylinder.  

Calculating activation energies

For the catalase-catalysed reaction, the relationship between reaction rate and concentration of hydrogen peroxide can be represented as rate = k[H 2 O 2 ] x [catalase] y , where k is the rate constant, the exponent,  x , is the order of the reaction with respect to hydrogen peroxide, and the exponent,  y , is the order of the reaction with respect to catalase. 

If we apply the method of initial rate, the concentrations of hydrogen peroxide and catalase can be assumed to be unchanged. (The initial rate of a reaction is the instantaneous rate determined just after the reaction begins,  ie   just after time = 0.)  

Thus, the above rate law may be replaced by initial rate = ck, where c is a constant. 

The variation of reaction rate with temperature is given by the Arrhenius equation, which in its integrated form is: 

k = Ae - E a / RT  

where A is a constant, the frequency factor,  E a  is the activation energy,  R  is the universal gas constant (8.314 J mol -1 K -1 ), and  T   is the absolute temperature. Since initial rate = ck, the initial rate can be found by: 

initial rate = cAe - E a / RT  

Taking logarithms: 

ln (initial rate) = lncA -  E a / RT

log (initial rate) = logcA -  E a /2.303 RT

Thus, we can plot log (initial rate) against 1/ T   to obtain a straight line. The slope is multiplied by -2.303 R  to get  E a . The rates can be expressed in volume of oxygen per second or in arbitrary units because the slope of the straight line will not be affected. 

Students' misconceptions

The initial students' plans revealed misconceptions that they had about their challenge:

  • They did not appreciate how to manipulate the Arrhenius equation. Some students thought that they could not determine the activation energy for a chemical reaction experimentally unless they first found the absolute values of the rate constants for the reactions. The rate constants at different temperatures can be calculated either from the rate law or from the integrated rate equation. However, because the rate law was not given in the problem and the concentration of catalase in lettuce is unknown, some students did not know how to proceed. In fact, the activation energy can be determined without knowing the rate law of a reaction (see Box 1). 
  • Temperatures were selected arbitrarily.  Some students forgot to check the optimum temperature for the enzymatic reaction. The optimum temperature for catalase is determined by the balance between the effect of temperature on the rate of the enzymatic reaction and its effect on the rate of inactivation of catalase. It also varies with the source and purity of catalase. The optimum temperature of catalase is  ca  40°C. 2  Students should therefore do preliminary experiments at two temperatures ( eg  40 and 45°C) to test the effect of temperature on lettuce catalase activity, and then select at least four different temperatures for the main investigation. The temperature intervals should be at least 5°C.  
  • They did not know what would constitute a fair test.  Some students believed that the activation energies for catalase-catalysed and iodide-catalysed reactions must be measured under the same experimental conditions, including the temperature range, and the concentrations of hydrogen peroxide and catalysts. In fact, activation energy is essentially independent of temperature and concentrations of reactants and catalysts. However, temperature and concentrations of reactants do affect the initial reaction rates. For each reaction, the students must do preliminary experiments to find suitable concentrations and volumes of hydrogen peroxide and catalyst solutions so that a measurable volume of oxygen is produced within a short period of time. 
  • Average rates were used.  Many students planned to measure the reaction rate by recording the oxygen produced after a period of time ( eg  two minutes) and then calculate the average rate. However, measurement of reaction rates is affected by numerous extraneous factors such as pH changes and the presence of side reactions. To minimise these factors, the method of initial rate should be used to determine activation energies. 

Students' results

After doing several pilot experiments, students finalised their procedures for measuring the rate of the catalase-catalysed reaction. Box 2 shows an example of the experimental procedure written by a group of students. They plotted the volume of oxygen evolved (cm 3 ) per second at the different temperatures, and calculated initial rates from the slopes of the straight lines drawn through the data points. Students used the Excel program to draw the straight lines on the plots. Although they used a crude sample of lettuce catalase and collected oxygen over water, the points in the Arrhenius plot (log (initial rate) versus 1/ T ×  10 3 ) lay satisfactorily on a straight line (R 2 = 0.94). They found that the activation energy for the catalase-catalysed reaction is 19 kJ mol -1 . (A reported value from the literature is 18kJ mol -1 .) 3  Note that teachers should be wary of comparing their students' results with the literature values because different sources of catalase affect the  E a values. For this catalase investigation, the  E a values obtained by students were 15-20 kJ mol -1 . 

The students used the same experimental set-up ( Fig 1 ) to measure the rate of the iodide-catalysed reaction, but they put 5 cm 3 of 0.1 M potassium iodide into the weighing bottle and 20 cm 3 of 0.22 M hydrogen peroxide into the conical flask. They did five experiments, at five different temperatures over the range 31-50°C, to calculate the initial rates. The students drew the straight lines based on the data points below 6 cm 3 of oxygen. They found that the activation energy is 55 kJ mol -1 (R 2 = 0.97), which is in good agreement with the literature value of 56 kJ mol -1 . 4 Again, the methods used by the students often give approximate  E a values only, ranging from 54-58 kJ mol -1 .

In general, activation energies for enzyme-catalysed reactions are within the range of 25- 63 kJ mol -1 . Thus, catalase is not only more efficient than inorganic catalysts such as iodide, but also an exceptionally efficient enzyme. A low value of activation energy implies high efficiency of molecular collisions to form products. The factor, e  - E a RT , in the Arrhenius equation is the proportion of molecules with energy equal to or greater than the activation energy. Since the activation energy for the uncatalysed decomposition of hydrogen peroxide is 75 kJ mol -1 , 5 the proportion at 25°C is equal to e -75000/(8.314 × 298) = 6.9 × 10 -14 . 

The rates of the iodide-catalysed and catalase-catalysed decomposition of hydrogen peroxide can be compared with that of the uncatalysed reaction. The relative rates are shown in Table 1. Lowering the activation energy from 75 to 55 kJ mol -1 increases the rate by a factor of 10 3 , but catalase is even more efficient than iodide by a factor of 10 6 . 

Image - Activation-Tab1

Achieving consistent results

Although the experimental procedure shown in Box 2 is crude, students can get consistent results if they keep the following points in mind: 

  • the ripeness of lettuce influences catalase activity and the amount of catalase varies from tissue to tissue. Extraction of catalase should only be done with fresh leaves;  
  • at room temperature, catalase in the extract dissociates and loses its activity rapidly. Keep the catalase extract in an ice-water bath, use the same catalase extract for all experiments, and try to complete the lab work within two hours; 
  • shake the reaction flask vigorously and continuously after the chemicals are mixed, otherwise less oxygen will be collected owing to the solubility of oxygen in water; 
  • low temperatures should be avoided, otherwise more oxygen gas will dissolve in the water;  
  • accuracy in maintaining the temperature of the water bath is essential for consistent results. Wait for at least 10 minutes to allow time for the chemicals to come to the same temperature as the water bath. Measure the temperature of the reaction mixture rather than the temperature of the water bath;
  • for safety, avoid hydrogen peroxide concentrations higher than 0.44 M ( ie  1.5 per cent). Note that concentrated hydrogen peroxide solution will inactivate catalase and oxidise iodide rapidly to produce iodine: 

(H 2 O 2 + 2I - + 2H + → I 2 + 2H 2 O); 

  • reading the measuring cylinder is harder than reading the stopwatch when the reaction proceeds rapidly. An alternative way to collect data would be to record the times necessary to produce two, three, four, and so on cm 3 of oxygen; 
  • the initial rate can be found by drawing the tangent to the volume versus time curve at  t = 0. However, this may not work because the first and second data points are usually unreliable owing to the difficulties associated with collecting the gases. Students should try to focus on the initial segment of the curve and then draw a best straight line. Researchers find that the error introduced by approximating the tangent is well within the normal limits for kinetic studies. 6

The catalase enzyme 

The kinetics of the catalase-catalysed decomposition of hydrogen peroxide is complicated and abnormal. Catalase is a redox enzyme responsible for the disproportionation reaction: 

2H 2 O 2  → 2H 2 O + O 2

Image - Feature - Cheung - main1

Source: © istockphoto

Fresh lettuce - a rich source of catalase

Catalase activity is particularly high in our liver, kidney and red blood cells. At temperatures higher than 40°C, the molecular and atomic motion disrupts the coiled globular structure of catalase, thus reducing its catalytic activity. The enzyme is said to be denatured. Catalase can also be inactivated by low-frequency ultrasound. Reseachers report that the activation energy for the thermal inactivation of beef liver catalase is 185 kJ mol -1 while the activation energy for the ultrasound inactivation of beef liver catalase is just 82 kJ mol -1 within the temperature range from 36- 55°C. 7  This implies that ultrasound can denature catalase more efficiently than heat. 

The catalase protein consists of four subunits. The active component of each subunit is a ferriproto-porphyrin group which is responsible for the disproportionation reaction. Researchers have discovered that an enzyme-substrate complex is involved, but the mechanism of the disproportionation reaction is still not fully understood. 8,9  The enzymatic reaction is thought to comprise two steps: 

[catalase-Fe(III)] + H 2 O 2   → H 2 O + [catalase-Fe(IV)=O] 

[catalase-Fe(IV)=O] + H 2 O 2  →   H 2 O + O 2  + [catalase-Fe(III)] 

Catalase is a very efficient catalyst. Each molecule of catalase can convert 4 × 10 7 H 2 O 2 molecules to products per second under optimal conditions. 10  In enzymatic reactions, the rate-determining step is normally the cleavage of the enzyme-substrate complex into enzyme and products. However, catalase shows abnormal kinetics - both steps are rate-determining. Consequently, the measured activation energy in this investigation is actually the resultant of the effects of temperature on two rate-determining processes instead of one.  

Dr Derek Cheung is associate professor in the department of curriculum and instruction at The Chinese University of Hong Kong, Shatin, Hong Kong (e-mail: [email protected]).

Acknowledgements

I would like to thank the Quality Education Fund for financial support of this project (grant code 2003/0750).

  • S. Mueller, H. D. Riedel and W. Stremmel,  Anal. Biochem ., 1997,   245 , 55.
  • D. R. Kimbrough, M. A. Magoun and M. Langfur,  J. Chem. Educ ., 1997,    74  (2), 210. 
  • I. W. Sizer,  J. Biolog. Chem ., 1944,   154 , 461. 
  • H. A. Liebhafsky and A. Mohammad,  J. Am. Chem. Soc ., 1933,   55 , 3977. 
  • E. A. Moelwyn-Hughes,  The chemical statics and kinetics of solutions . London: Academic, 1971. 
  • J. Casado, M. A. Lopez-Quintela and F. M. Lorenzo-Barral,  J. Chem. Educ ., 1986,   63  (5), 450. 
  • M. V. Potapovich, A. N. Eremin and D. I. Metelitza,  Appl. Biochem. and Microbiol ., 2003,   39 , 140. 
  • P. Jones,  J. Biolog. Chem ., 2001,    276 , 13791. 
  • S. G . Kalka  et al ,  J. Am. Chem. Soc ., 2001,   123 , 9665. 
  • J. W. Hill, S. J. Baum and R. J. Scott-Ennis,  Chemistry and life . Upper Saddle River, NJ: Prentice Hall, 2000. 
  • Maths skills
  • Misconceptions
  • Practical skills and safety
  • Problem-based learning
  • Rates of reaction
  • Secondary education

Related articles

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

Small bowls of different fermented foods - gherkins, sour cream, kimchi, kombucha, saurkraut

Why fermented foods are good for your gut – and your teaching

2024-04-15T05:30:00Z By Emma Davies

From kimchi to kefir, tuck into the complex chemistry of fermentation and its health potential

1 Reader's comment

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

A small man standing on a triangle

Improve students’ understanding with Johnstone’s triangle

2024-08-19T07:12:00Z By Duncan Short and Colin McGill

Make abstract ideas concrete by linking macroscopic, submicroscopic and symbolic representations

STEM person of the week cards spread out showing photos of real scientists, a short description of their job and their personal attributes that help them in that role

Here’s how to showcase science careers students can aspire to

2024-08-05T05:22:00Z By Carol Davenport

Get learners to see a future in science by highlighting attributes and skills

A tablet computer showing different states of matter and a beaker of water

Enhance students’ learning and development with digital resources

2024-06-17T05:02:00Z By David Paterson

Tips and a model to improve learners’ understanding and develop vital skills using digital learning

  • Contributors
  • Print issue
  • Email alerts

Site powered by Webvision Cloud

  • Architecture and Design
  • Asian and Pacific Studies
  • Business and Economics
  • Classical and Ancient Near Eastern Studies
  • Computer Sciences
  • Cultural Studies
  • Engineering
  • General Interest
  • Geosciences
  • Industrial Chemistry
  • Islamic and Middle Eastern Studies
  • Jewish Studies
  • Library and Information Science, Book Studies
  • Life Sciences
  • Linguistics and Semiotics
  • Literary Studies
  • Materials Sciences
  • Mathematics
  • Social Sciences
  • Sports and Recreation
  • Theology and Religion
  • Publish your article
  • The role of authors
  • Promoting your article
  • Abstracting & indexing
  • Publishing Ethics
  • Why publish with De Gruyter
  • How to publish with De Gruyter
  • Our book series
  • Our subject areas
  • Your digital product at De Gruyter
  • Contribute to our reference works
  • Product information
  • Tools & resources
  • Product Information
  • Promotional Materials
  • Orders and Inquiries
  • FAQ for Library Suppliers and Book Sellers
  • Repository Policy
  • Free access policy
  • Open Access agreements
  • Database portals
  • For Authors
  • Customer service
  • People + Culture
  • Journal Management
  • How to join us
  • Working at De Gruyter
  • Mission & Vision
  • De Gruyter Foundation
  • De Gruyter Ebound
  • Our Responsibility
  • Partner publishers

redox reaction and catalase activity experiment

Your purchase has been completed. Your documents are now available to view.

Catalase, a remarkable enzyme: targeting the oldest antioxidant enzyme to find a new cancer treatment approach

This review is centered on the antioxidant enzyme catalase and will present different aspects of this particular protein. Among them: historical discovery, biological functions, types of catalases and recent data with regard to molecular mechanisms regulating its expression. The main goal is to understand the biological consequences of chronic exposure of cells to hydrogen peroxide leading to cellular adaptation. Such issues are of the utmost importance with potential therapeutic extrapolation for various pathologies. Catalase is a key enzyme in the metabolism of H 2 O 2 and reactive nitrogen species, and its expression and localization is markedly altered in tumors. The molecular mechanisms regulating the expression of catalase, the oldest known and first discovered antioxidant enzyme, are not completely elucidated. As cancer cells are characterized by an increased production of reactive oxygen species (ROS) and a rather altered expression of antioxidant enzymes, these characteristics represent an advantage in terms of cell proliferation. Meanwhile, they render cancer cells particularly sensitive to an oxidant insult. In this context, targeting the redox status of cancer cells by modulating catalase expression is emerging as a novel approach to potentiate chemotherapy.

Historical view

The origin of catalase (EC 1.11.1.6), the enzyme that metabolizes H 2 O 2 but also reacts with a multitude of other substrates, can be traced back to the 19th century (Figure 1 ) when Thénard discovered H 2 O 2 and suspected that its tissue degradation in living organisms was the result of a ‘special’ substance activity ( Thénard, 1811 ). Schönbein showed that a ‘ferment’ can detoxify H 2 O 2 ( Schönbein, 1863 ) and later on Loew gave the name ‘catalase’ to the enzyme converting H 2 O 2 into water and oxygen ( Loew, 1900 ). Loew found the presence of this enzyme in many organisms ranging from plants to mammals. In the 1920s, Warburg and coworkers demonstrated, by using cyanide as enzyme inhibitor, that the active site of catalase contains an iron atom ( Warburg, 1923 ). Furthermore, Stern showed that the hemin group of the enzyme can react with compounds such as cyanides, sulfides, fluorides and he also showed that the enzyme active group had a ferric complex identical to the protoporphyrin found in the hemoglobin of red blood cells ( Stern, 1937 ). Thereafter, Sumner and Dounce (1937) purified and crystallized bovine catalase.

Figure 1: Time line of catalase main discoveries and findings.Modified from Sies (2017).

Time line of catalase main discoveries and findings.

Modified from Sies (2017) .

The advances in biochemistry facilitated further elucidation about the mechanisms of the ‘catalase’ enzymatic reaction. Thus, the work carried out by Chance’s lab led to the discovery of the formation of the ‘Compound I’ that occurred during the reaction between catalase and the first molecule of H 2 O 2 . A few years later, he also discovered the ‘compounds II and III’ ( Chance, 1948 , 1949 ; Chance et al., 1952 ). In the 1960s, it was elucidated the role of key residues in the active site of the enzyme, such as the distal histidine, and their importance for the stabilization of the tertiary structure was discussed by Nakatani (1961) . In 1970, catalase Compound I was identified in intact eukaryotic cells, proving the existence of H 2 O 2 in normal aerobic metabolism ( Sies and Chance, 1970 ). Kirkman and Gaetani reported that NADPH was the enzyme cofactor bound to catalase ( 1984 ), which was further confirmed following X-ray analysis of catalase structures ( Fita and Rossmann, 1985a ). Finally, recombinant phage clones containing the human catalase gene were isolated and characterized ( Quan et al., 1986 ) and the use of molecular biology techniques have led to major advances in the understanding of catalase regulation mechanisms and the role that this enzyme could play in numerous biological processes.

In this context, Nenoi et al. (2001) reported that in the upstream catalase promoter, there is a region containing CCAAT and GGGCGG boxes where transcription factors like Nuclear factor Y (NF-Y) and Specificity protein 1 (Sp1) can bind to the catalase promoter regulating the transcriptional activation of the human catalase gene. Recently, Glorieux et al. (2016a) identified a new regulatory region in the human catalase promoter, in which chromatin remodeling is required to regulate catalase expression by retinoic acid receptor alpha (RARα) and JunB transcription factors. This novel regulatory mechanism is involved during cancer cells adaptation to chronic exposure to H 2 O 2 and may have therapeutic consequences for various diseases and metabolic disorders.

Types of catalases

With the increasing number of complete sequences available, distinct homologies were detected and catalases came to be classified in three groups based on their structure and function. The first and the second group contain heme-containing enzymes, namely typical or true catalases and catalase-peroxidases, whereas the third group contains (non-heme) manganese catalases ( Zamocky and Koller, 1999 ; Zamocky et al., 2008 , 2010 ).

Typical catalases

The members of this largest group are found in aerobically respiring organisms. In contrast, in anaerobic bacteria, catalases proteins are generally not expressed. Most of these catalases are homotetramers, between 200 and 340 kDa in size and contain four prosthetic groups. In the majority of true catalases, a ferric protoporphyrin IX was found in the active center (namely heme b , similar to the prosthetic group of human hemoglobin). Some variants do exist such as ‘heme d ’ groups which can reside in some typical catalases.

Following phylogenetic analyses the typical catalases can also be divided into three main clades. Clade 1 contains bacterial, algal and plant catalases with small-subunit size (55–69 kDa) using heme b as the prosthetic group. Clade 2 regroups bacterial and fungal catalases and has a large subunit size (75–84 kDa), with heme d as the prosthetic group and an additional ‘flavodoxin-like’ domain. Clade 3 is the most abundant subfamily; catalases from this subgroup are found in archaebacterial, fungi, protists, plants and animals. The human catalase belongs to this clade and is characterized by a small subunit (62 kDa), with heme b as its prosthetic group and NADPH as cofactor.

Catalase-peroxidases

These proteins have been found in fungi, archeobacteria and bacteria. Their molecular weight varies between 120 and 340 kDa and they are generally homodimers. The catalase activity (degrading hydrogen peroxide) is less efficient than in typical catalases but catalase-peroxidases have a better affinity for their substrate H 2 O 2 . Catalase-peroxidases are also significantly more sensitive than typical catalases to inactivation by pH and temperature. The well-known horseradish peroxidase, currently employed in immunoblotting experiments, is one example of catalase-peroxidase.

Manganese catalases

These enzymes have been found exclusively in bacteria. Manganese catalases utilize two manganese ions in the active site, they can form oligomeric structures measuring between 170 and 210 kDa and have no significant homology with either typical catalases or catalase-peroxidases. The catalytic reaction is completely different to other types of catalases. The dimanganese core is equally stable Mn 2+ –Mn 2+ or Mn 3+ –Mn 3+ . Like typical catalases, the catalase reaction occurs in two-step.

1st step: H 2 O 2 + Mn 2 + ​ ​ − ​ Mn 2 + (2H + ) → Mn 3+ ​ − ​ Mn 3+ + 2 H 2 O 2nd step: H 2 O 2 + Mn 3 + ​ − ​ Mn 3 + → Mn 2 + ​ − ​ Mn 2 + (2H + ) + O 2 Sum of reactions: H 2 O 2 + H 2 O 2 → 2 H 2 O + O 2

Structure of human catalase

Human catalase contains four identical subunits of 62 kDa, each subunit containing four distinct domains and one prosthetic heme group ( Nagem et al., 1999 ; Zamocky and Koller, 1999 ; Putnam et al., 2000 ). The four domains include: (1) a N-terminal arm which contains a distal histidine, an essential amino acid for the catalase reaction; (2) a β-barrel domain that contains eight β-barrels arranged in an antiparallel fashion with six α-helical insertions, conferring the hydrophobic core of the protein necessary for the tri-dimensional structure of the enzyme; (3) a connection domain which contains the tyrosine residue that binds the heme group; and finally (4) an α-helical domains, which is important for NADPH binding.

Although amino acid sequences do not have high identities between all typical catalases, the tridimensional structure is highly conserved. The tertiary structure of the β-barrel domain, the connection domain and the zone neighboring the distal histidine are highly conserved. The α-helical domain is moderately conserved between species and some typical catalases do not bind the cofactor NADPH.

The investigation of inhibitory mechanisms by which cyanide and 3-amino-1,2,4-triazole (ATA) inhibit human catalase allowed to understand the enzyme activity ( Putnam et al., 2000 ). Cyanide nitrogen blocks heme access to other potential ligands. It interacts with the distal histidine and an asparagine residue suggesting that it competes with hydrogen peroxide for heme binding. Meanwhile, ATA interacts with the distal histidine leading to an adduct formation and thereby blocks the catalase reaction.

Mechanism of ‘catalase’ reaction

During the enzymatic reaction leading to H 2 O 2 destruction, catalase is first oxidized to a hypervalent iron intermediate, known as compound I ( Cpd I ), which is then reduced back to the resting state by a second H 2 O 2 molecule.

1st reaction: Enz (Porf-Fe III ) + H 2 O 2 →   C p d I  (Porf • + -Fe IV ​ ​ ​ = ​ O) + H 2 O 2nd reaction:  C p d I  (Porf • + -Fe IV ​ ​ = ​ O) + H 2 O 2 →   Enz (Porf-Fe III ) + H 2 O + O 2 Sum of reactions: H 2 O 2 + H 2 O 2 → 2 H 2 O + O 2

The first reaction is characterized by the oxidation of the heme protein by a single H 2 O 2 molecule leading to the formation of Cpd I , an oxoferryl porphyrin cation radical ( Jones and Dunford, 2005 ). Once Cpd I is formed, it reacts rapidly with a second molecule of H 2 O 2 to generate H 2 O and O 2 in a two-electron redox process. This second reaction is particularly efficient in some catalases compared to other heme proteins such as myoglobin ( Matsui et al., 1999 ). Labeling studies have shown that both H 2 O and O 2 molecules are formed from the same molecule of H 2 O 2 ( Vlasits et al., 2007 ). Fita and Rossmann (1985b) proposed that two molecules of H 2 O 2 were sequentially transferred to the oxoferryl group of Cpd I , where the distal histidine residue plays a role of acid-base catalyst. Although mutation of distal histidine suppresses the ability to form Cpd I ( Nakatani, 1961 ), some authors claimed that reaction occurs as a direct mechanism and histidine does not play a crucial role in catalysis ( Kato et al., 2004 ).

In the presence of one-electron donors (such as phenols, ferrocyanide, salicylic acid, NO, superoxide anions) and low H 2 O 2 concentrations, Cpd I may undergo a one-electron reduction towards the inactive compound II ( Cpd II ) intermediate, which transforms back to the resting state through another one-electron reduction step (reviewed in Bauer, 2015 ). Both Cpd I and Cpd II are generally described as an oxoferryl-heme species ( Rovira, 2005 ). In the case of Cpd I , the porphyrin bears a cation radical (O=Fe IV -heme˙ + ), while Cpd II lacks the porphyrin cation radical (O=Fe IV -heme). Thus, Cpd II is best described as a hydroxoferryl bond (HO-Fe IV -heme) instead of the traditional oxoferryl species, consistent with the fact that a proton is released upon conversion of Cpd I to Cpd II :

C p d I + AH (one-electron donor) →   C p d I I (Por-Fe IV ​ =   O and Por-Fe IV -OH)

At higher H 2 O 2 concentrations, NADPH prevents the generation of Cpd II by participating in a two-electron reduction process ( Kirkman et al., 1999 ). In the presence of another one-electron donor, Cpd II will return to a resting state. But in presence of a H 2 O 2 molecule, Cpd II will be transformed back to compound III ( Cpd III ), an inactive intermediate ( Gabdoulline et al., 2003 ). In this intermediate state, iron is at an oxyferrous state (O 2 -Fe II -heme). Then, Cpd III goes back to a resting state or leads to the inactivation of the catalase.

C p d I I + AH → Enz (Por-Fe III ) resting state C p d I I + H 2 O 2 → C p d I I I  (Por-Fe II -O 2 ) C p d I I I → Enz (Por-Fe III ) or inactivation

Cellular and tissue catalase distribution

Regarding catalase localization within the cell, it should be noted that catalase is mainly located in peroxisomes because it contains a sequence signal recognized by some peroxisome receptors. Contrary to mitochondria, proteins located within peroxisomes are all of nuclear origin and should be imported. Indeed, it is generally accepted that catalase monomers are imported into peroxisomes where tetramerization and heme addition occurs ( Lazarow and De Duve, 1973 ). The apoprotein (monomer) enters into the peroxisome by a peroxisome-targeting signal sequence (PTS) present on the carboxy-terminal tail of catalase. The most common targeting signal is the SKL (serine-lysine-leucine) but catalase is characterized by a different signal, namely the KANL (lysine-alanine-asparagine-leucine) signal ( Purdue and Lazarow, 1996 ). Proteins bearing these signals are recognized by the PTS1 receptor, called PEX5p for humans. Some diseases related to defects in peroxisome biogenesis, such as Zellweger syndrome, are characterized by mutations in the PEX5p receptor and cellular H 2 O 2 overproduction due to a catalase default import in peroxisomes ( Wanders et al., 1984 ). This syndrome is most commonly called ‘the syndrome of empty peroxisome’. It has been shown that overexpressing receptor mutants in PEX5-deficient CHO cells, drastically reduce the import of proteins such as catalase ( Shimozawa et al., 1999 ). Some authors have observed that catalase with the SKL signal and not KANL had a better import capacity and could be transported into the peroxisome even in the case of mutations in the PTS1 receptor ( Koepke et al., 2007 ). It has been shown that PEX5p recognizes catalase which is already folded and interacts with other receptors such as PEX13p for proper import into the peroxisome ( Otera and Fujiki, 2012 ). Studies are underway to validate, through clinical trials, whether catalase SKL has a therapeutic potential for diseases with peroxisome biogenesis disorders. Recently, it has been reported that PEX19p, an essential protein for peroxisome biogenesis, interacts with Valosin-containing protein (VCP) and regulates the catalase cytoplasmic localization, a potential feedback mechanism modulating H 2 O 2 levels ( Murakami et al., 2013 ).

Interestingly, the existence of a cytosolic catalase, in its active tetrameric conformation, has been reported ( Middelkoop et al., 1993 ) with varied percentage from one cell type to another, and different function to peroxisome catalase. Indeed, catalase may bind cytosolic proteins such as Grb2 and SHP2, to protect them from potential oxidative damage ( Yano et al., 2004a , b ). These proteins are linked to the membrane by a pleckstrin homology (PH) domain, so it is common to find catalase in fractions including membrane proteins and membrane-associated proteins. In this context, it has been shown that catalase can be localized at the cytoplasmic membrane, specifically at the surface of cancer cells ( Bauer, 2012 ). This locally high expression of catalase on the membrane of tumor cells is in line with the findings by Deichman’s group that showed that tumor progression in vivo is dependent on increased resistance towards exogenous H 2 O 2 ( Deichman, 2000 , 2002 ). Furthermore, localized expression of catalase on the membrane of tumor cells is not in disagreement with the finding of lower total catalase concentration in malignant cells, as the membrane comprises a minority of the total cellular material.

These observations open the way for new anti-cancer therapies targeting catalase with specific antibodies ( Bauer, 2012 ; Bauer and Motz, 2016 ) or exogenous singlet oxygen ( Riethmüller et al., 2015 ; Bauer and Graves, 2016 ) to induce apoptosis by reactivation of intercellular HOCl and/or NO/peroxynitrite signaling after catalase inhibition or inactivation. Furthermore, the modulation of the intracellular NO concentration has been shown to lead to the generation of cell-derived singlet oxygen that inactivates tumor cell protective catalase and reactivates intercellular ROS/RNS-mediated apoptosis-inducing signaling ( Bauer, 2015 ; Scheit and Bauer, 2015 ).

In addition to classical intracellular catalase and membrane-associated catalase of tumor cells ( Heinzelmann and Bauer, 2010 ; Böhm et al., 2015 ), the release of soluble catalase from tumor cells has also been reported ( Sandstrom and Buttke, 1993 ; Moran et al., 2002 ; Böhm et al., 2015 ). This soluble extracellular catalase was protective for the tumor cells.

Finally, catalase has been also localized in the mitochondria of rat cardiomyocytes ( Radi et al., 1991 ).

The human catalase is expressed in every organ and the highest levels of activity are measured in the liver, kidney and red blood cells ( Winternitz and Meloy, 1908 ). In erythrocytes, a high production of H 2 O 2 is generated due to oxygen transport and catalase is responsible for more than 50% of the H 2 O 2 turnover ( Mueller et al., 1997 ).

Biological functions of catalase and related diseases

The first function assigned to catalase is the dismutation of H 2 O 2 into oxygen and water without consummation of endogenous reducing equivalents, an important role in cell defense against oxidative damage by H 2 O 2 . To note that H 2 O 2 is not only toxic by its ability to form other ROS, like hydroxyl radical through the Fenton reaction ( Fenton, 1894 ) but as was nicely recently reviewed by Sies, H 2 O 2 acting as a second messenger is involved in many biological processes including changes of morphology, proliferation, signaling (i.e. NF-κB), apoptosis, etc ( Sies, 2017 ). In addition to its dominant ‘catalatic’ activity (decomposition of H 2 O 2 ), catalase can also act in its peroxidatic mode, i.e. decomposition of small substrates such as methanol, formate, azide, hydroperoxides ( Sies, 1974 ; Chance et al., 1979 ; Johansson and Borg, 1988 ), and in case of ethanol, it is also capable of oxidize it to acetaldehyde contributing to its liver metabolism ( Keilin and Hartree, 1945 ; Thurman et al., 1972 ; Oshino et al., 1973 ). It has also been reported that catalase may decompose peroxynitrite ( Gebicka and Didik, 2009 ; Heinzelmann and Bauer, 2010 ), oxidize nitric oxide to nitrite ( Wink and Mitchell, 1998 ; Brunelli et al., 2001 ). A discrete balance between oxidation of NO by compound I of catalase and inhibition of catalase by NO through formation of a CAT-Fe III NO complex has been reported ( Brown, 1995 ). Catalase also exhibits low oxidase activity (O 2 -dependent oxidation of organic substrates) ( Vetrano et al., 2005 ). Thus, catalase may also have additional roles such as the detoxification or activation of toxic and anti-tumor compounds. For instance, catalase has been detected in mouse oocytes most likely to protect the genome from oxidative damage during meiotic maturation ( Park et al., 2016 ).

Within this framework, several studies have shown a change in catalase expression in cancer cells became resistant to chemotherapies ( Akman et al., 1990 ; Kim et al., 2001 ; Kalinina et al., 2006 ). Thus, a potential role of catalase during the acquisition of cancer cell resistance to chemotherapeutic agents was explored by overexpressing the human enzyme in MCF-7 cells, a human derived breast cancer cell line ( Glorieux et al., 2011 ). No particular resistance against conventional chemotherapies like doxorubicin, cisplatin and paclitaxel was observed in cells overexpressing catalase but they were more resistant to the pro-oxidant effect induced by an H 2 O 2 -generating system ( Glorieux et al., 2011 ).

In addition, catalase mitochondria overexpression in mice enables an increase of lifespan by 20% ( Schriner et al., 2005 ). In such animals, the development of mitochondrial deletions was reduced and heart disease and the onset of cataracts were delayed.

Regarding catalase down-regulation, no particular sensitivity was observed as catalase-deficient mice are viable and fertile ( Ho et al., 2004 ). They develop normally with a normal hematological profile, but after trauma the mitochondria shows defects in the oxidative phosphorylation. Note that humans may also be deficient in catalase, a condition known as acatalasemia that is characterized by a low catalase rate, but it is still rare and usually benign ( Goth et al., 2004 ).

In this context, there are benign polymorphisms of the catalase gene for which no change of catalase expression or activity was detected ( Goth et al., 2004 ). They include single nucleotide substitutions in the promoter, 5′ untranslated region, intron 1, exon 1, exon 9 and 10 ( Goth et al., 2004 ). To note that the catalase gene encodes one single protein of 526 amino acids and the single locus has been mapped to chromosome 11p13 ( Wieacker et al., 1980 ). The length of the catalase gene is 34 kb, it contains 12 introns and 13 exons generating a mRNA of 2286 bp ( Quan et al., 1986 ).

Conversely, some catalase mutations provoke changes in either catalase expression or activity and may be associated with some diseases. For instance, a higher transcriptional activity in two human cancer cell lines, namely HepG2 and K562 cells, was observed in the case of common functional C-T substitution polymorphism in the promoter region (−262) of the human catalase gene ( Forsberg et al., 2001 ) but no mutations have been detected in the coding sequence, to our knowledge, in patients suffering from cancer. Recent studies have focused on the associations of catalase polymorphisms with various types of cancer but many inconsistent results about the relationship between the catalase gene polymorphism and cancer risk were reported. Recently, two meta-analyses pointed out a correlation exists between this polymorphism C-262T and prostate cancer ( Liu et al., 2016 ; Wang et al., 2016 ).

Different catalase mutations in patients can cause decreased catalase activity leading to increased H 2 O 2 concentrations in the blood and tissues. Table 1 shows the polymorphism of the catalase gene in patients suffering from hypocatalasemia (about 50% of catalase activities) or acatalasemia (less than 10% of catalase activities). Depending on the mutations, such patients may be subject to an increased risk of type 2 diabetes, vitiligo and increased blood pressure ( Goth et al., 2004 ). When the mutations are located in exons as in Japanese and Hungarian acatalasemia, a truncated or mutated catalase is synthetized and functionally less active. Takahara was the first to describe acatalasemia and this pathology is often benign but Japanese patients can suffer from oral gangrenes and esophageal ulcerations (Takahara’s disease), probably promoted by H 2 O 2 generated by phagocytic cells and bacterial actions ( Takahara, 1952 ).

Catalase gene polymorphisms in patients suffering from hypocatalasemia or acatalasemia.

MutationsPositionRegionAssociated disease
G insertionCodon 48Exon 2Diabetes type 2
T→ACodon 53Exon 2Diabetes type 2
G→CCodon 66Exon 2Diabetes type 2
C→T−773 PromoterHypertension
T→CCodon 388Exon 9Vitiligo
G→A5 Intron 4Japanese hypocatalasemia (A)
T deletionCodon 134Exon 4Japanese hypocatalasemia (B)
GA insertionCodon 67Exon 2Hungarian hypocatalasemia (A)
G insertionCodon 58Exon 2Hungarian hypocatalasemia (B)
G→T5 Intron 7Hungarian hypocatalasemia (C)
G→ACodon 354Exon 9Hungarian hypocatalasemia (D)

This table was adapted from Goth et al. (2004) .

a Position from the ATG translation start.

b Position from the first nucleotide in the intron sequence.

Decreased activity of catalase has also been observed in various genetic alterations, for example, loss of alleles (i.e. loss of heterozygosity) of the catalase gene in non-small-cell lung cancer cells ( Ludwig et al., 1991 ; Fong et al., 1994 ; Shipman et al., 1998 ) or deletion of chromosome 11p, as has been observed in children affected by Wilms’ tumor, aniridia, gonadoblastoma and retardation (WAGR) syndrome ( Dufier et al., 1981 ; Gregoire et al., 1983 ; Barletta et al., 1985 ; Michalopoulos et al., 1985 ; van Heyningen et al., 1985 ).

Regulation of catalase expression in cancer cells

It is generally accepted that the cellular maintenance of redox homeostasis is controlled by a complex network of antioxidant enzymes (i.e. superoxide dismutases and glutathione peroxidases) whose expression is under the fine-tuning control of the Keap1-Nrf2 signaling pathway ( Menegon et al., 2016 ). Nevertheless, the molecular mechanisms regulating the expression of catalase – the oldest known and first discovered antioxidant enzyme – are independent of this pathway and not totally elucidated. Therefore, the fine-tuning regulation of this enzyme should be prior elucidated in order to find a new approach to modulate the antioxidant status in cancer cells.

Interestingly, despite the existence of diverse protection mechanisms against oxidant injuries, a consensus emerged in the scientific literature about an alteration of redox homeostasis within tumor cells. Indeed, they produce large amounts of ROS that are involved in the maintenance of genetic instability favoring cancer cell proliferation. Meanwhile, altered expression levels of catalase have been reported in cancer tissues as compared to their normal counterparts. Thus, as compared to normal tissues of the same origin, some authors reported an increased catalase expression in tumors ( Sander et al., 2003 ; Hwang et al., 2007 ; Rainis et al., 2007 ), whereas other studies showed a catalase down-regulation ( Marklund et al., 1982 ; Baker et al., 1997 ; Lauer et al., 1999 ; Chung-man et al., 2001 ; Cullen et al., 2003 ; Kwei et al., 2004 ), indicating that cancer cells are frequently more sensitive to an oxidative stress. For instance, we have reported an important decrease of catalase activity in different cancer cell lines, as shown in Table 2 ( Verrax et al., 2009 ; Beck et al., 2011a ; Glorieux et al., 2011 ). Briefly, catalase levels can vary after short treatments to H 2 O 2 ( Rohrdanz and Kahl, 1998 ; Rohrdanz et al., 2001 ; Sen et al., 2003 ) and catalase expression is modified in cancer cell lines rendered resistant to chronic exposures to H 2 O 2 ( Kasugai and Yamada, 1992 ; Nenoi et al., 2001 ) or certain chemotherapeutic compounds such as doxorubicin ( Ramu et al., 1984 ; Akman et al., 1990 ; Kim et al., 2001 ; Kalinina et al., 2006 ). Although mechanisms controlling catalase expression have been partially elucidated, the decreased catalase expression in cancer cells still remains an unanswered question.

Catalase enzyme activity in cells from diver origins.

Type of cellsNormal originCancer originReferences
Mouse hepatocytes96.45±6.3211.12±4.43
Human leukocytes44.55±1.8016.36±3.60
Human mammary epithelial cells13.14±3.04 5.44±0.88

Catalase activity (U/mg proteins) in normal and cancer cell lines. a TLT mouse hepatocarcinoma cells; b K562 human CML cells; c 250MK cells; and d MCF-7 human mammary cells. Data were analyzed by unpaired t -test. e p -Value≤0.01; f p -value≤0.001.

The regulation of catalase expression in cancer cells is a complex process because different levels of regulation are thought to be involved. In a recent review, we discussed the different mechanisms playing a potential role in the regulation of its expression in both healthy and tumor cells ( Glorieux et al., 2015 ). They include transcriptional regulation, represented by the activity of transcription factors that induce or repress the transcriptional activity of catalase promoters, post-transcriptional regulation (mRNA stability) and post-translational modification (phosphorylation and ubiquitination of the protein). In addition, epigenetic (DNA methylation, modifications of histones) changes or genetic alterations can also be involved playing a role in governing proper levels of catalase activity in these cells.

Regarding transcription it should be noted that the catalase gene has all the characteristics of a housekeeping gene (no TATA box, no INR sequence, high GC content in promoter) and a core promoter which is highly conserved among species ( Quan et al., 1986 ; Nakashima et al., 1990 ; Reimer et al., 1994 ). In this core promoter, the presence of DNA binding sites for transcription factors like NF-Y and Sp1 has an essential role in the positive regulation of catalase expression ( Nenoi et al., 2001 ). Additional transcription factors have also been involved in this regulatory process. In fact, there is strong evidence that the protein Akt/PKB in the PI3K signaling pathway plays a major role in the expression of catalase by modulating the activity of FoxO3a ( Turdi et al., 2007 ; Venkatesan et al., 2007 ; Akca et al., 2013 ). Therefore, targeting PI3K/Akt/mTOR ( LoPiccolo et al., 2008 ; Rodon et al., 2013 ) may be an efficient way to increase the expression of catalase in tumors and inhibit tumor cell growth.

In this last decade, other transcription factors (PPARγ, Oct-1, etc.) as well as genetic, epigenetic and post-transcriptional processes are emerging as crucial contributors to the negative regulation of catalase expression ( Glorieux et al., 2015 ). Specifically, we investigated the transcriptional regulatory mechanism controlling catalase expression in human mammary cell lines. To this end, we have made a human breast MCF-7 cancer cell line resistant to oxidative stress, the so-called Resox cells. These cells show decreased ROS basal levels and an increased activity of some antioxidant enzymes, notably catalase ( Dejeans et al., 2012 ; Glorieux et al., 2015 , 2016b ). A novel promoter region, responsible for the regulation of catalase expression, was identified at −1518/−1226 locus in Resox cells ( Glorieux et al., 2016a ). The AP-1 family member JunB and retinoic acid receptor alpha (RARα) mediate catalase transcriptional activation and repression, respectively, by controlling chromatin remodeling through a histone deacetylases-dependent mechanism ( Glorieux et al., 2016a ). Indeed, RARα and JunB act in collaboration with transcription factors or other proteins on either a closed- or an open-promoter chromatin status regulating the expression of catalase in breast cancer cells (Figure 2 ). Thus, cancer adaptation to oxidative stress, regulated by transcriptional factors through chromatin remodeling appears as a new mechanism to target cancer cells.

Figure 2: Hypothetical mechanism of catalase regulation in breast cancer cell lines.A human breast MCF-7 cancer cell line resistant to oxidative stress was generated (Resox cells) by chronically exposing MCF-7 cells with an H2O2-generating system (Dejeans et al., 2012). These cells show increased expression of catalase. A novel promoter region, responsible for the regulation of catalase expression, was identified at −1518/−1226 locus in Resox cells (Glorieux et al., 2016a). The AP-1 family member JunB and retinoic acid receptor alpha (RARα) mediate catalase transcriptional activation and repression, respectively, by controlling chromatin remodeling through a histone deacetylases (HDAC)-dependent mechanism. (HAT: Histobe acetyltransferase). Adapted from Glorieux et al. (2016a).

Hypothetical mechanism of catalase regulation in breast cancer cell lines.

A human breast MCF-7 cancer cell line resistant to oxidative stress was generated (Resox cells) by chronically exposing MCF-7 cells with an H 2 O 2 -generating system ( Dejeans et al., 2012 ). These cells show increased expression of catalase. A novel promoter region, responsible for the regulation of catalase expression, was identified at −1518/−1226 locus in Resox cells ( Glorieux et al., 2016a ). The AP-1 family member JunB and retinoic acid receptor alpha (RARα) mediate catalase transcriptional activation and repression, respectively, by controlling chromatin remodeling through a histone deacetylases (HDAC)-dependent mechanism. (HAT: Histobe acetyltransferase). Adapted from Glorieux et al. (2016a) .

Oxidative stress-based therapies against cancer

As previously mentioned, cancer cells are generally deficient in antioxidant enzymes, thus, any increase in ROS levels would be a menace to the precarious redox balance of cancer cells making them vulnerable to an additional oxidative stress. Given that weakness, the loss of redox homeostasis represents an interesting target for research and development of new molecules with antitumor activity and numerous drugs are currently being clinically evaluated. Therefore, several strategies have been developed looking for the disruption of tumor cell redox homeostasis and a subsequent cancer cell death ( Demizu et al., 2008 ; Trachootham et al., 2009 ; Verrax et al., 2009 ).

One of these strategies is the use of ROS-generating compounds such as arsenic trioxide (ATO), currently employed against promyelocytic leukemia ( Valenzuela et al., 2014 ; Lo-Coco et al., 2016 ), or doxorubicin, an anthracycline used for the treatment of several types of cancer (i.e. breast cancer). Indeed, the impairment of mitochondrial function due to increased levels of superoxide anion is supposed to be the main mechanism of both chemotherapeutic drugs ( Thayer, 1977 ; Pelicano et al., 2003 ). A decrease in antioxidant levels has also been proposed. For instance, glutathione (GSH) levels may be decreased either by its direct binding with phenylethylisothiocyanate (PEITC) and sulphoraphane ( Trachootham et al., 2009 ) or by inhibition of γ-glutamylcysteine synthetase activity (a key enzyme involved in GSH synthesis) as done by Buthionine sulfoximine ( Griffith and Meister, 1979 ). The development of specific inhibitors of thioredoxin and thioredoxin reductases was also carried out. Inhibitors of thioredoxin, such as PX-12, were shown to have potent antitumor activities ( Welsh et al., 2003 ; Baker et al., 2006 ). Conversely, the overexpression of Trx1 is correlated with resistance to anti-cancer drugs ( Baker et al., 2006 ; Kaimul et al. 2007 ).

As quinones display redox cycling abilities thus generating ROS ( Kappus and Sies, 1981 ), the association of menadione (a naphthoquinone derivative) and ascorbate (Figure 3 ), was employed to trigger tumor cell death ( Verrax et al., 2011b ). This association (Asc/Men) has been shown to have potent in vitro and in vivo antitumor activities enhancing as well the therapeutic effect of currently used anticancer drugs ( Taper et al., 1987 ; Buc Calderon et al., 2002 ; Verrax et al., 2003 ). We hypothesized that H 2 O 2 , issued from the redox cycling, is the oxidant species responsible for antitumor effects observed both in vitro and in vivo ( Verrax et al., 2006 , 2007 ; Verrax and Buc Calderon, 2009 ). The induced oxidative stress provokes cell necrosis by a wide variety of processes including ATP depletion ( Verrax et al., 2004 , 2005 , 2011a ), disruption of Ca 2+ homeostasis ( Dejeans et al., 2010 ) and loss of HSP90 chaperone activity ( Beck et al., 2009 , 2011b , 2012 ). Based on the vulnerability of tumor cells to an oxidative stress, we have induced the alteration of their intracellular redox homeostasis as a new strategy in the research and development of new antitumor drugs ( Benites et al., 2010 ; Arenas et al., 2013 ; Valderrama et al., 2015 ). A similar approach has been recently developed by using the SnFe 2 O 4 nanocrystals, a heterogeneous Fenton catalyst, which once internalized into the cancer cells, convert H 2 O 2 into hydroxyl radicals inducing apoptotic cell death. In normal cells, the oxidative injury induced by SnFe 2 O 4 is prevented by catalase ( Lee et al., 2017 ).

Figure 3: The quinone redox cycling hypothesis.The biological activity showed by menadione mainly relies upon its ability to accept one electron from ascorbate to form a semiquinone radical. When the semiquinone is reduced back to its quinone form, superoxide anion is produced which by dismutation generates hydrogen peroxide.

The quinone redox cycling hypothesis.

The biological activity showed by menadione mainly relies upon its ability to accept one electron from ascorbate to form a semiquinone radical. When the semiquinone is reduced back to its quinone form, superoxide anion is produced which by dismutation generates hydrogen peroxide.

As ATO increases the levels of ROS ( Valenzuela et al., 2014 ) and Asc/Men increases the cytotoxicity of chemotherapeutic drugs ( Taper et al., 1987 ), we formulated the following hypothesis: Asc/Men potentiates ATO-mediated cytotoxicity (Figure 4 ). ATO decreases catalase expression most likely by activating Akt pathway and/or inducing RARα, meanwhile Asc/Men generates H 2 O 2 . Despite that Resox cells display high antioxidant defenses ( Dejeans et al., 2012 ; Glorieux et al., 2015 , 2016b ), an enhanced cell death was observed when they were exposed to a mixture containing sublethal doses of Asc/Men and ATO. This is likely due to a decreased transcriptional activity of the human catalase −1518/+16 promoter caused by ATO resulting in less catalase protein (Glorieux et al., 2017, unpublished results).

Figure 4: Arsenic trioxide (ATO) decreases catalase expression in breast cancer cells and sensitizes them to pro-oxidant drugs.

Arsenic trioxide (ATO) decreases catalase expression in breast cancer cells and sensitizes them to pro-oxidant drugs.

Conclusions

Under stress conditions, the antioxidant enzyme catalase plays a major role by detoxifying H 2 O 2 . As consequences, a change of its activity or expression will lead to pathological processes as Zellweger syndrome, acatalasemia or WAGR syndrome. The subcellular localization of catalase is mainly peroxisomal but a shuttle between this organelle and cytoplasm exists and may be involved in the protection of key cellular elements (i.e. proteins, chromosomes) against an oxidative damage.

Our group and others demonstrated that catalase expression is also altered in cancer cells, most likely to favor cell proliferation by inducing genetic instability and activation of oncogenes. The regulation of catalase expression appears to be mainly controlled at transcriptional levels although other mechanisms may also be involved. In addition of transcription factors like Sp1 and NF-Y, JunB and RARα transcription factors are crucial regulators in breast cancer cells by recruiting proteins involved in transcriptional complexes and chromatin remodeling.

Therefore, catalase can be a future therapeutic target in the context of cancer by using pro-oxidant approaches.

Acknowledgments

The authors thank Professor Helmut Sies for the splendid discussion and his precious input. This project was funded by FNRS-Televie Grant (grant n° 7.4575.12F).

Akca, H., Demiray, A., Aslan, M., Acikbas, I., and Tokgun, O. (2013). Tumour suppressor PTEN enhanced enzyme activity of GPx, SOD and catalase by suppression of PI3K/AKT pathway in non-small cell lung cancer cell lines. J. Enzyme Inhib. Med. Chem. 28 , 539–544. 10.3109/14756366.2011.654114 Search in Google Scholar

Akman, S.A., Forrest, G., Chu, F.F., Esworthy, R.S., and Doroshow, J.H. (1990). Antioxidant and xenobiotic-metabolizing enzyme gene expression in doxorubicin-resistant MCF-7 breast cancer cells. Cancer Res. 50 , 1397–1402. Search in Google Scholar

Arenas, P., Pena, A., Rios, D., Benites, J., Muccioli, G.G., Buc Calderon, P., and Valderrama, J.A. (2013). Echo-friendly synthesis and antiproliferative evaluation of oxygen substituted diaryl ketones. Molecules 18 , 9818–9832. 10.3390/molecules18089818 Search in Google Scholar

Baker, A.M., Oberley, L.W., and Cohen, M.B. (1997). Expression of antioxidant enzymes in human prostatic adenocarcinoma. Prostate 32 , 229–233. 10.1002/(SICI)1097-0045(19970901)32:4<229::AID-PROS1>3.0.CO;2-E Search in Google Scholar

Baker, A.F., Dragovich, T., Tate, W.R., Ramanathan, R.K., Roe, D., Hsu, C.H., Kirkpatrick, D.L., and Powis, G. (2006). The antitumor thioredoxin-1 inhibitor PX-12 (1-methylpropyl 2-imidazolyl disulfide) decreases thioredoxin-1 and VEGF levels in cancer patient plasma. J Lab Clin Med. 147 , 83–90. 10.1016/j.lab.2005.09.001 Search in Google Scholar

Barletta, C., Castello, M.A., Ferrante, E., Mavelli, I., Clerico, A., Ciriolo, M.R., and Vignetti, P. (1985). 11p13 deletion and reduced RBC catalase in a patient with aniridia, glaucoma and bilateral Wilms’ tumor. Tumori 71 , 119–121. 10.1177/030089168507100205 Search in Google Scholar

Bauer, G. (2012). Tumor cell-protective catalase as a novel target for rational therapeutic approaches based on specific intercellular ROS signaling. Anticancer Res. 32 , 2599–2624. Search in Google Scholar

Bauer, G. (2015). Increasing the endogenous NO level causes catalase inactivation and reactivation of intercellular apoptosis signaling specifically in tumor cells. Redox Biol. 6 , 353–371. 10.1016/j.redox.2015.07.017 Search in Google Scholar

Bauer, G. and Graves, D.B. (2016). Mechanisms of selective antitumor action of cold atmospheric plasma-derived reactive oxygen and nitrogen species. Plasma Process Polym. 13 , 1157–1178. 10.1002/ppap.201600089 Search in Google Scholar

Bauer, G. and Motz, M. (2016). The antitumor effect of single-domain antibodies directed towards membrane-associated catalase and superoxide dismutase. Anticancer Res. 36 , 5945–5956. 10.21873/anticanres.11182 Search in Google Scholar

Beck, R., Verrax, J., Gonze, T., Zappone, M., Curi Pedrosa, R., Taper, H., Feron, O., and Buc Calderon, P. (2009). Hsp90 cleavage by an oxidative stress leads to Bcr-Abl degradation and leukemia cell death. Biochem. Pharmacol. 77 , 375–383. 10.1016/j.bcp.2008.10.019 Search in Google Scholar

Beck, R., Pedrosa, R.C., Dejeans, N., Glorieux, C., Leveque, P., Gallez, B., Taper, H., Eeckhoudt, S., Knoops, L., Buc Calderon, P., et al. (2011a). Ascorbate/menadione-induced oxidative stress kills cancer cells that express normal or mutated forms of the oncogenic protein Bcr-Abl. An in vitro and in vivo mechanistic study. Invest New Drugs 29 , 891–900. 10.1007/s10637-010-9441-3 Search in Google Scholar

Beck, R., Dejeans, N., Glorieux, C., Curi Pedrosa, R., Vasquez, D., Valderrama, J.A., Buc Calderon, P., and Verrax, J. (2011b). Molecular chaperone Hsp90 as a target for oxidant-based anticancer therapies. Curr. Med. Chem. 18 , 2816–2825. 10.2174/092986711796011256 Search in Google Scholar

Beck, R., Dejeans, N., Glorieux, C., Creton, M., Delaive, E., Dieu, M., Raes, M., Leveque, P., Gallez, B., Depuydt, M., et al. (2012). Hsp90 is cleaved by reactive oxygen species at a highly conserved N-terminal amino acid motif. PLoS One 7 , e40795. 10.1371/journal.pone.0040795 Search in Google Scholar

Benites, J., Valderrama, J.A., Bettega, K., Curi Pedrosa, R., Buc Calderon, P., and Verrax, J. (2010). Biological evaluation of donor-acceptor aminonaphthoquinones as antitumor agents. Eur. J. Med. Chem. 45 , 6052–6057. 10.1016/j.ejmech.2010.10.006 Search in Google Scholar

Böhm, B., Heinzelmann, S., Motz, M., and Bauer, G. (2015). Extracellular localization of catalase is associated with the transformed state of malignant cells. Biol Chem. 396 , 1339–1356. 10.1515/hsz-2014-0234 Search in Google Scholar

Brown, G.C. (1995). Reversible binding and inhibition of catalase by nitric oxide. Eur J Biochem. 232 , 188–191. 10.1111/j.1432-1033.1995.tb20798.x Search in Google Scholar

Brunelli, L., Yermilov, V., and Beckman, J.S. (2001). Modulation of catalase peroxidatic and catalatic activity by nitric oxide. Free Radic. Biol. Med. 30 , 709–714. 10.1016/S0891-5849(00)00512-8 Search in Google Scholar

Buc Calderon, P., Cadrobbi, J., Marquez, C., Hong-Ngo, N., Jamison, J.M., Gilloteaux, J., Summers, J.L., and Taper, H.S. (2002). Potential therapeutic application of the association of vitamins C and K3 in cancer treatment. Curr. Med. Chem. 9 , 2271–2285. 10.2174/0929867023368674 Search in Google Scholar

Chance, B. (1948). The enzyme-substrate compounds of catalase and peroxides. Nature 161 , 914–917. 10.1038/161914a0 Search in Google Scholar

Chance, B. (1949). The primary and secondary compounds of catalase and methyl or ethyl hydrogen peroxide; kinetics and activity. J. Biol. Chem. 179 , 1341–1369. 10.1016/S0021-9258(18)56799-5 Search in Google Scholar

Chance, B., Greenstein, D.S., and Roughton, F.J. (1952). The mechanism of catalase action. I. Steady-state analysis. Arch. Biochem. Biophys. 37 , 301–321. 10.1016/0003-9861(52)90194-X Search in Google Scholar

Chance, B., Sies, H., and Boveris, A. (1979). Hydroperoxide metabolism in mammalian organs. Physiol Rev. 59 , 527–605. 10.1152/physrev.1979.59.3.527 Search in Google Scholar

Chung-man, H.J., Zheng, S., Comhair, S.A., Farver, C., and Erzurum, S.C. (2001). Differential expression of manganese superoxide dismutase and catalase in lung cancer. Cancer Res. 61 , 8578–8585. Search in Google Scholar

Cullen, J.J., Mitros, F.A., and Oberley, L.W. (2003). Expression of antioxidant enzymes in diseases of the human pancreas: another link between chronic pancreatitis and pancreatic cancer. Pancreas 26 , 23–27. 10.1097/00006676-200301000-00005 Search in Google Scholar

Deichman, G.I. (2000). Natural selection and early changes of phenotype of tumor cells in vivo: acquisition of new defense mechanisms. Biochemistry (Mosc). 65 , 78–94. Search in Google Scholar

Deichman, G.I. (2002). Early phenotypic changes of in vitro transformed cells during in vivo progression: possible role of the host innate immunity. Semin. Cancer Biol. 12 , 317–326. 10.1016/S1044-579X(02)00018-4 Search in Google Scholar

Dejeans, N., Tajeddine, N., Beck, R., Verrax, J., Taper, H., Gailly, P., and Buc Calderon, P. (2010). Endoplasmic reticulum calcium release potentiates the ER stress and cell death caused by an oxidative stress in MCF-7 cells. Biochem. Pharmacol. 79 , 1221–1230. 10.1016/j.bcp.2009.12.009 Search in Google Scholar

Dejeans, N., Guenin, S., Glorieux, C., Beck, R., Sid, B., Rousseau, R., Bisig, B., Delvenne, P., Buc Calderon, P., and Verrax, J. (2012). Overexpression of GRP94 in breast cancer cells resistant to oxidative stress promotes high levels of cancer cell proliferation and migration: implications for tumor recurrence. Free Radic. Biol. Med. 52 , 993–1002. 10.1016/j.freeradbiomed.2011.12.019 Search in Google Scholar

Demizu, Y., Sasaki, R., Trachootham, D., Pelicano, H., Colacino, J.A., Liu, J., and Huang, P. (2008). Alterations of cellular redox state during NNK-induced malignant transformation and resistance to radiation. Antiox. Redox Signal. 10 , 951–961. 10.1089/ars.2007.1871 Search in Google Scholar

Dufier, J.L., Phug, L.H., Schmelck, P., Fekete, C., de Grouchy, G.J., Turleau, C., and Haye, C. (1981). [Intercalary deletion of the short arm of chromosome 11: aniridia, glaucoma, staturoponderal and mental retardation, sexual ambiguity, gonadoblastoma and catalase deficiency]. Bull. Soc. Ophtalmol. Fr. 81 , 747–749. Search in Google Scholar

Fenton, H.J.H. (1894). Oxidation of tartaric acid in presence of iron. J. Chem. Soc. Trans. 65 , 899–911. 10.1039/CT8946500899 Search in Google Scholar

Fita, I. and Rossmann, M.G. (1985a). The NADPH binding site on beef liver catalase. Proc. Natl. Acad. Sci. USA 82 , 1604–1608. 10.1073/pnas.82.6.1604 Search in Google Scholar

Fita, I. and Rossmann, M.G. (1985b). The active center of catalase. J. Mol. Biol. 185 , 21–37. 10.1016/0022-2836(85)90180-9 Search in Google Scholar

Fong, K.M., Zimmerman, P.V., and Smith, P.J. (1994). Correlation of loss of heterozygosity at 11p with tumour progression and survival in non-small cell lung cancer. Genes Chromosomes Cancer 10 , 183–189. 10.1002/gcc.2870100306 Search in Google Scholar PubMed

Forsberg, L., Lyrenas, L., de Faire, U., and Morgenstern, R. (2001). A common functional C-T substitution polymorphism in the promoter region of the human catalase gene influences transcription factor binding, reporter gene transcription and is correlated to blood catalase levels. Free Radic. Biol. Med. 30 , 500–505. 10.1016/S0891-5849(00)00487-1 Search in Google Scholar

Gabdoulline, R.R., Kummer, U., Olsen, L.F., and Wade, R.C. (2003). Concerted simulations reveal how peroxidase compound III formation results in cellular oscillations. Biophys. J. 85 , 1421–1428. 10.1016/S0006-3495(03)74574-3 Search in Google Scholar

Gebicka, L. and Didik, J. (2009). Catalytic scavenging of peroxynitrite by catalase. J. Inorg. Biochem. 103 , 1375–1379. 10.1016/j.jinorgbio.2009.07.011 Search in Google Scholar

Glorieux, C., Dejeans, N., Sid, B., Beck, R., Buc Calderon, P., and Verrax, J. (2011). Catalase overexpression in mammary cancer cells leads to a less aggressive phenotype and an altered response to chemotherapy. Biochem. Pharmacol. 82 , 1384–1390. 10.1016/j.bcp.2011.06.007 Search in Google Scholar

Glorieux, C., Zamocky, M., Sandoval, C.J.M., Verrax, J., and Buc Calderon, P. (2015). Regulation of catalase expression in healthy and cancer cells. Free Radic. Biol. Med. 87 , 84–97. 10.1016/j.freeradbiomed.2015.06.017 Search in Google Scholar

Glorieux, C., Sandoval, J.M., Fattaccioli, A., Dejeans, N., Garbe, J.C., Dieu, M., Verrax, J., Renard, P., Huang, P., and Buc Calderon, P. (2016a). Chromatin remodeling regulates catalase expression during cancer cells adaptation to chronic oxidative stress. Free Radic. Biol. Med. 99 , 436–450. 10.1016/j.freeradbiomed.2016.08.031 Search in Google Scholar

Glorieux, C., Sandoval, J.M., Dejeans, N., Ameye, G., Poirel, H.A., Verrax, J., and Buc Calderon P. (2016b). Overexpression of NAD(P)H:quinone oxidoreductase 1 (NQO1) and genomic gain of the NQO1 locus modulates breast cancer cell sensitivity to quinones. Life Sci. 145 , 57–65. 10.1016/j.lfs.2015.12.017 Search in Google Scholar

Goth, L., Rass, P., and Pay, A. (2004). Catalase enzyme mutations and their association with diseases. Mol. Diagn. 8 , 141–149. 10.1007/BF03260057 Search in Google Scholar

Gregoire, M.J., Pernot, C., Himont, F., Pierson, M., and Gilgenkrantz, S. (1983). [Chromosome 11 and cancer]. J.Genet.Hum. 31 , 31–36. Search in Google Scholar

Griffith, O.W. and Meister, A. (1979). Potent and specific inhibition of glutathione synthesis by buthionine sulfoximine (S-n-butyl homocysteine sulfoximine). J. Biol. Chem. 254 , 7558–7560. 10.1016/S0021-9258(18)35980-5 Search in Google Scholar

Heinzelmann, S. and Bauer, G. (2010). Multiple protective functions of catalase against intercellular apoptosis-inducing ROS signalling of human tumor cells, Biol. Chem. 391 , 675–693. Search in Google Scholar

Ho, Y.S., Xiong, Y., Ma, W., Spector, A., and Ho, D.S. (2004). Mice lacking catalase develop normally but show differential sensitivity to oxidant tissue injury. J. Biol. Chem. 279 , 32804–32812. 10.1074/jbc.M404800200 Search in Google Scholar PubMed

Hwang, T.S., Choi, H.K., and Han, H.S. (2007). Differential expression of manganese superoxide dismutase, copper/zinc superoxide dismutase, and catalase in gastric adenocarcinoma and normal gastric mucosa. Eur. J. Surg. Oncol. 33 , 474–479. 10.1016/j.ejso.2006.10.024 Search in Google Scholar

Johansson, L.H. and Borg, L.A. (1988). A spectrophotometric method for determination of catalase activity in small tissue samples, Anal. Biochem. 174 , 331–336. Search in Google Scholar

Jones, P. and Dunford, H.B. (2005). The mechanism of Compound I formation revisited. J. Inorg. Biochem. 99 , 2292–2298. 10.1016/j.jinorgbio.2005.08.009 Search in Google Scholar

Kaimul, A.M., Nakamura, H., Masutani, H., and Yodoi, J. (2007). Thioredoxin and thioredoxin-binding protein-2 in cancer and metabolic syndrome. Free Radic. Biol. Med. 43 , 861–868. 10.1016/j.freeradbiomed.2007.05.032 Search in Google Scholar

Kalinina, E.V., Chernov, N.N., Saprin, A.N., Kotova, Y.N., Andreev, Y.A., Solomka, V.S., and Scherbak, N.P. (2006). Changes in expression of genes encoding antioxidant enzymes, heme oxygenase-1, Bcl-2, and Bcl-xl and in level of reactive oxygen species in tumor cells resistant to doxorubicin. Biochemistry (Mosc.) 71 , 1200–1206. 10.1134/S0006297906110058 Search in Google Scholar

Kappus, H. and Sies, H. (1981). Toxic drug effects associated with oxygen metabolism: redox cycling and lipid peroxidation. Experientia 37 , 1233–1241. 10.1007/BF01948335 Search in Google Scholar

Kasugai, I. and Yamada, M. (1992). High production of catalase in hydrogen peroxide-resistant human leukemia HL-60 cell lines. Leuk. Res. 16 , 173–179. 10.1016/0145-2126(92)90129-U Search in Google Scholar

Kato, S., Ueno, T., Fukuzumi, S., and Watanabe, Y. (2004). Catalase reaction by myoglobin mutants and native catalase: mechanistic investigation by kinetic isotope effect. J. Biol. Chem. 279 , 52376–52381. 10.1074/jbc.M403532200 Search in Google Scholar PubMed

Keilin, D. and Hartree, E.F. (1945). Properties of catalase. Catalysis of coupled oxidation of alcohols. Biochem. J. 39 , 293–301. 10.1042/bj0390293 Search in Google Scholar

Kim, H.S., Lee, T.B., and Choi, C.H. (2001). Down-regulation of catalase gene expression in the doxorubicin-resistant AML subline AML-2/DX100. Biochem. Biophys. Res. Commun. 281 , 109–114. 10.1006/bbrc.2001.4324 Search in Google Scholar PubMed

Kirkman, H.N. and Gaetani G.F. (1984). Catalase: a tetrameric enzyme with four tightly bound molecules of NADPH. Proc.Natl. Acad. Sci. USA 81 , 4343–4347. 10.1073/pnas.81.14.4343 Search in Google Scholar PubMed PubMed Central

Kirkman, H.N., Rolfo, M., Ferraris, A.M., and Gaetani, G.F. (1999). Mechanisms of protection of catalase by NADPH. Kinetics and stoichiometry. J. Biol. Chem. 274 , 13908–13914. 10.1074/jbc.274.20.13908 Search in Google Scholar PubMed

Koepke, J.I., Nakrieko, K.A., Wood, C.S., Boucher, K.K., Terlecky, L.J., Walton, P.A., and Terlecky, S.R. (2007). Restoration of peroxisomal catalase import in a model of human cellular aging. Traffic 8 , 1590–1600. 10.1111/j.1600-0854.2007.00633.x Search in Google Scholar PubMed

Kwei, K.A., Finch, J.S., Thompson, E.J., and Bowden, G.T. (2004). Transcriptional repression of catalase in mouse skin tumor progression. Neoplasia 6 , 440–448. 10.1593/neo.04127 Search in Google Scholar PubMed PubMed Central

Lauer, C., Volkl, A., Riedl, S., Fahimi, H.D., and Beier, K. (1999). Impairment of peroxisomal biogenesis in human colon carcinoma. Carcinogenesis 20 , 985–989. 10.1093/carcin/20.6.985 Search in Google Scholar PubMed

Lazarow, P.B. and De Duve, C. (1973). The synthesis and turnover of rat liver peroxisomes. V. Intracellular pathway of catalase synthesis. J. Cell Biol. 59 , 507–524. 10.1083/jcb.59.2.507 Search in Google Scholar PubMed PubMed Central

Lee, K.T., Lu, Y.J., Mi, F.L., Burnouf, T., Wei, Y.T., Chiu, S.C., Chuang, E.Y., and Lu, S.Y. (2017). Catalase-modulated heterogeneous Fenton reaction for selective cancer cell eradication: SnFe 2 O 4 nanocrystals as an effective reagent for treating lung cancer cells. ACS Appl. Mater. Interfaces 9 , 1273–1279. 10.1021/acsami.6b13529 Search in Google Scholar PubMed

Liu, K., Liu, X., Wang, M., Wang, X., Kang, H., Lin, S., Yang, P., Dai, C., Xu, P., Li, S., et al. (2016). Two common functional catalase gene polymorphisms (rs1001179 and rs794316) and cancer susceptibility: evidence from 14,942 cancer cases and 43,285 controls. Oncotarget 7 , 62954–62965. 10.18632/oncotarget.10617 Search in Google Scholar PubMed PubMed Central

Lo-Coco, F., Cicconi, L., and Breccia, M. (2016). Current standard treatment of adult acute promyelocytic leukaemia. Br. J. Haematol. 172 , 841–854. 10.1111/bjh.13890 Search in Google Scholar PubMed

Loew, O. (1900). A new enzyme of general occurrence in organisms. Science 11 , 701–702. 10.1126/science.11.279.701 Search in Google Scholar PubMed

LoPiccolo, J., Blumenthal, G.M., Bernstein, W.B., and Dennis, P.A. (2008). Targeting the PI3K/Akt/mTOR pathway: effective combinations and clinical considerations. Drug Resist. Updat. 11 , 32–50. 10.1016/j.drup.2007.11.003 Search in Google Scholar PubMed PubMed Central

Ludwig, C.U., Raefle, G., Dalquen, P., Stulz, P., Stahel, R., and Obrecht, J.P. (1991). Allelic loss on the short arm of chromosome 11 in non-small-cell lung cancer. Int. J. Cancer 49 , 661–665. 10.1002/ijc.2910490506 Search in Google Scholar

Marklund, S.L., Westman, N.G., Lundgren, E., and Roos, G. (1982). Copper- and zinc-containing superoxide dismutase, manganese-containing superoxide dismutase, catalase, and glutathione peroxidase in normal and neoplastic human cell lines and normal human tissues. Cancer Res. 42 , 1955–1961. Search in Google Scholar

Matsui, T., Ozaki, S., Liong, E., Phillips, G.N., and Watanabe, Jr., Y. (1999). Effects of the location of distal histidine in the reaction of myoglobin with hydrogen peroxide. J. Biol. Chem. 274 , 2838–2844. 10.1074/jbc.274.5.2838 Search in Google Scholar

Menegon, S., Columbano, A., and Giordano, S. (2016). The dual roles of Nrf2 in cancer. Trends Mol. Med. 22 , 578–593. 10.1016/j.molmed.2016.05.002 Search in Google Scholar

Michalopoulos, E.E., Bevilacqua, P.J., Stokoe, N., Powers, V.E., Willard, H.F., and Lewis, W.H. (1985). Molecular analysis of gene deletion in aniridia-Wilms tumor association. Hum. Genet. 70 , 157–162. 10.1007/BF00273074 Search in Google Scholar

Middelkoop, E., Wiemer, E.A., Schoenmaker, D.E., Strijland, A., and Tager, J.M. (1993). Topology of catalase assembly in human skin fibroblasts. Biochim. Biophys. Acta 1220 , 15–20. 10.1016/0167-4889(93)90091-3 Search in Google Scholar

Moran, E.C., Kamiguti, A.S., Cawley, J.C., and Pettitt, A.R. (2002). Cytoprotective antioxidant activity of serum albumin and autocrine catalase in chronic lymphocytic leukaemia. Br. J. Haematol. 116 , 316–328. 10.1046/j.1365-2141.2002.03280.x Search in Google Scholar

Mueller, S., Riedel, H.D., and Stremmel, W. (1997). Direct evidence for catalase as the predominant H 2 O 2 -removing enzyme in human erythrocytes. Blood 90 , 4973–4978. 10.1182/blood.V90.12.4973 Search in Google Scholar

Murakami, K., Ichinohe, Y., Koike, M., Sasaoka, N., Iemura, S., Natsume, T., and Kakizuka, A. (2013). VCP is an integral component of a novel feedback mechanism that controls intracellular localization of catalase and H 2 O 2 Levels. PLoS One 8 , e56012. 10.1371/journal.pone.0056012 Search in Google Scholar

Nagem, R.A., Martins, E.A., Goncalves, V.M., Aparicio, R., and Polikarpov, I. (1999). Crystallization and preliminary X-ray diffraction studies of human catalase. Acta Crystallogr. D Biol. Crystallogr. 55 , 1614–1615. 10.1107/S0907444999009695 Search in Google Scholar

Nakashima, H., Yamamoto, M., Goto, K., Osumi, T., Hashimoto, T., and Endo, H. (1990). Isolation and characterization of the rat catalase-encoding gene. Gene 89 , 295. 10.1016/0378-1119(90)90021-I Search in Google Scholar

Nakatani, M. (1961). Studies on histidine residues in hemeproteins related to their activities. V. Photooxidation of catalase in the presence of methylene blue. J. Biochem. 49 , 98–102. 10.1093/oxfordjournals.jbchem.a127278 Search in Google Scholar

Nenoi, N., Ichimura, S., Mita, K., Yukawa,O., and Cartwright, I.L. (2001). Regulation of the catalase gene promoter by Sp1, CCAAT-recognizing factors, and a WT1/Egr-related factor in hydrogen peroxide-resistant HP100 cells. Cancer Res. 61 , 5885–5894. Search in Google Scholar

Oshino, N., Oshino, R., and Chance, B. (1973). The characteristics of the ‘peroxidatic’ reaction of catalase in ethanol oxidation. Biochem J. 131 , 555–563. 10.1042/bj1310555 Search in Google Scholar

Otera, H. and Fujiki, Y. (2012). Pex5p imports folded tetrameric catalase by interaction with Pex13p. Traffic 13 , 1364–1377. 10.1111/j.1600-0854.2012.01391.x Search in Google Scholar

Park, Y.S., You, S.Y., Cho, S., Jeon, H.J., Lee, S., Cho, D.H., Kim, J.S., and Oh, J.S. (2016). Eccentric localization of catalase to protect chromosomes from oxidative damages during meiotic maturation in mouse oocytes. Histochem. Cell Biol. 146 , 281–288. 10.1007/s00418-016-1446-3 Search in Google Scholar

Pelicano, H., Feng, L., Zhou, Y., Carew, J.S., Hileman, E.O., Plunkett, W., Keating, M.J., and Huang, P. (2003). Inhibition of mitochondrial respiration: a novel strategy to enhance drug-induced apoptosis in human leukemia cells by a reactive oxygen species-mediated mechanism. J. Biol. Chem. 278 , 37832–37839. 10.1074/jbc.M301546200 Search in Google Scholar

Purdue, P.E. and Lazarow, P.B. (1996). Targeting of human catalase to peroxisomes is dependent upon a novel COOH-terminal peroxisomal targeting sequence. J. Cell Biol. 134 , 849–862. 10.1083/jcb.134.4.849 Search in Google Scholar

Putnam, C.D., Arvai, A.S., Bourne, Y., and Tainer, J.A. (2000). Active and inhibited human catalase structures: ligand and NADPH binding and catalytic mechanism. J. Mol. Biol. 296 , 295–309. 10.1006/jmbi.1999.3458 Search in Google Scholar

Quan, F., Korneluk, R.G., Tropak, M.B., and Gravel, R.A. (1986). Isolation and characterization of the human catalase gene. Nucleic Acids Res. 14 , 5321–5335. 10.1093/nar/14.13.5321 Search in Google Scholar

Radi, R., Turrens, J.F., Chang, L.Y., Bush, K.M., Crapo, J.D., and Freeman, B.A. (1991). Detection of catalase in rat heart mitochondria. J. Biol. Chem. 266 , 22028–22034. 10.1016/S0021-9258(18)54740-2 Search in Google Scholar

Rainis, T., Maor, I., Lanir, A., Shnizer, S., and Lavy, A. (2007). Enhanced oxidative stress and leucocyte activation in neoplastic tissues of the colon. Dig. Dis. Sci. 52 , 526–530. 10.1007/s10620-006-9177-2 Search in Google Scholar PubMed

Ramu, A., Cohen, L., and Glaubiger, D. (1984). Oxygen radical detoxification enzymes in doxorubicin-sensitive and -resistant P388 murine leukemia cells. Cancer Res. 44 , 1976–1980. Search in Google Scholar

Reimer, D.L., Bailley, J., and Singh, S.M. (1994). Complete cDNA and 5′ genomic sequences and multilevel regulation of the mouse catalase gene. Genomics 21 , 325–336. 10.1006/geno.1994.1273 Search in Google Scholar

Rodon, J., Dienstmann, R., Serra, V., and Tabernero, J. (2013). Development of PI3K inhibitors: lessons learned from early clinical trials. Nat. Rev. Clin. Oncol. 10 , 143–153. 10.1038/nrclinonc.2013.10 Search in Google Scholar

Rohrdanz, E. and Kahl, R. (1998). Alterations of antioxidant enzyme expression in response to hydrogen peroxide. Free Radic. Biol. Med. 24 , 27–38. 10.1016/S0891-5849(97)00159-7 Search in Google Scholar

Rohrdanz, E., Schmuck, G., Ohler, S., and Kahl, R. (2001). The influence of oxidative stress on catalase and MnSOD gene transcription in astrocytes. Brain Res. 900 , 128–136. 10.1016/S0006-8993(01)02277-6 Search in Google Scholar

Rovira, C. (2005). Structure, protonation state and dynamics of catalase compound II. Chemphyschem. 6 , 1820–1826. 10.1002/cphc.200400633 Search in Google Scholar PubMed

Riethmüller, M., Burger. N., and Bauer, G. (2015). Singlet oxygen treatment of tumor cells triggers extracellular singlet oxygen generation, catalase inactivation and reactivation of intercellular apoptosis-inducing signaling. Redox Biol. 6 , 157–168. 10.1016/j.redox.2015.07.006 Search in Google Scholar PubMed PubMed Central

Sander, C.S., Hamm, F., Elsner, P., and Thiele, J.J. (2003). Oxidative stress in malignant melanoma and non-melanoma skin cancer. Br. J. Dermatol. 148 , 913–922. 10.1046/j.1365-2133.2003.05303.x Search in Google Scholar PubMed

Sandstrom, P.A. and Buttke, P.M. (1993). Autocrine production of extracellular catalase prevents apoptosis of the human CEM T-cell line in serum-free medium. Proc. Natl. Acad. Sci. USA 90 , 4708–4712. 10.1073/pnas.90.10.4708 Search in Google Scholar PubMed PubMed Central

Scheit, K. and Bauer, G. (2015). Direct and indirect inactivation of tumor cell protective catalase by salicylic acid and anthocyanidins reactivates intercellular ROS signaling and allows for synergistic effects. Carcinogenesis 36 , 400–411. 10.1093/carcin/bgv010 Search in Google Scholar PubMed

Schönbein, C.F. (1863). Chemische Mittheilungen. Med. Chir. Trans. 89 , 1–38. Search in Google Scholar

Schriner, S.E., Linford, N.J., Martin, G.M., Treuting, P., Ogburn, C.E., Emond, M., Coskun, P.E., Ladiges, W., Wolf, N., Van Remmen, H., et al. (2005). Extension of murine life span by overexpression of catalase targeted to mitochondria. Science 308 , 1909–1911. 10.1126/science.1106653 Search in Google Scholar PubMed

Sen, P., Mukherjee, S., Bhaumik, G., Das, P., Ganguly, S., Choudhury, N., and Raha, S. (2003). Enhancement of catalase activity by repetitive low-grade H 2 O 2 exposures protects fibroblasts from subsequent stress-induced apoptosis. Mutat. Res. 529 , 87–94. 10.1016/S0027-5107(03)00106-4 Search in Google Scholar

Shimozawa, N., Zhang, Z., Suzuki, Y., Imamura, A., Tsukamoto, T., Osumi, T., Fujiki, Y., Orii, T., Barth, P.G., Wanders, R.J., et al. (1999). Functional heterogeneity of C-terminal peroxisome targeting signal 1 in PEX5-defective patients. Biochem. Biophys. Res. Commun. 262 , 504–508. 10.1006/bbrc.1999.1232 Search in Google Scholar

Shipman, R., Schraml, P., Colombi, M., and Ludwig, C.U. (1998). Allelic deletion at chromosome 11p13 defines a tumour suppressor region between the catalase gene and D11S935 in human non-small cell lung carcinoma. Int. J. Oncol. 12 , 107–111. 10.3892/ijo.12.1.107 Search in Google Scholar

Sies, H. (1974). Biochemistry of the peroxisome in the liver cell. Angew. Chem. Int. Ed. Engl. 13 , 706–718. 10.1002/anie.197407061 Search in Google Scholar

Sies, H. (2017). Hydrogen peroxide as a central redox signaling molecule in physiological oxidative stress: oxidative eustress. Redox Biol. 11 , 613–619. 10.1016/j.redox.2016.12.035 Search in Google Scholar

Sies, H. and Chance, B. (1970). The steady state level of catalase compound I in isolated hemoglobin-free perfused rat liver. FEBS Lett. 11 , 172–176. 10.1016/0014-5793(70)80521-X Search in Google Scholar

Stern, K.G. (1937). Spectroscopy of catalase. J. Gen. Physiol. 20 , 631–648. 10.1085/jgp.20.4.631 Search in Google Scholar

Sumner, J.B. and Dounce, A.L. (1937). Crystalline catalase. Science 85 , 366–367. 10.1126/science.85.2206.366 Search in Google Scholar

Takahara, S. (1952). Progressive oral gangrene probably due to lack of catalase in the blood (acatalasaemia); report of nine cases. Lancet 2 , 1101–1104. 10.1016/S0140-6736(52)90939-2 Search in Google Scholar

Taper, H.S., de Gerlache, J., Lans, M., and Roberfroid, M. (1987). Non-toxic potentiation of cancer chemotherapy by combined C and K3 vitamin pre-treatment. Int. J. Cancer 40 , 575–579. 10.1002/ijc.2910400424 Search in Google Scholar PubMed

Thayer, W.S. (1977). Adriamycin stimulated superoxide formation in submitochondrial particles. Chem. Biol. Interact. 19 , 265–278. 10.1016/0009-2797(77)90050-3 Search in Google Scholar

Thénard, L.J. (1811). Observations sur des nouvelles combinaisons entre l’oxigène et divers acides. Ann. Chim. Phys. 8 , 306–312. Search in Google Scholar

Thurman, R.G., Ley, H.G., and Scholz, R. (1972). Hepatic microsomal ethanol oxidation. Hydrogen peroxide formation and the role of catalase. Eur. J. Biochem. 25 , 420–430. 10.1111/j.1432-1033.1972.tb01711.x Search in Google Scholar

Trachootham, D., Alexandre, J., and Huang, P. (2009). Targeting cancer cells by ROS-mediated mechanisms: a radical therapeutic approach? Nat. Rev. Drug Discov. 8 , 579–591. 10.1038/nrd2803 Search in Google Scholar

Turdi, S., Li, Q., Lopez, F.L., and Ren, J. (2007). Catalase alleviates cardiomyocyte dysfunction in diabetes: role of Akt, Forkhead transcriptional factor and silent information regulator 2. Life Sci. 81 , 895–905. 10.1016/j.lfs.2007.07.029 Search in Google Scholar

Valderrama, J.A., Rios, D., Muccioli, G.G., Buc Calderon, P., Brito, I., and Benites, J. (2015). Hetero-annulation reaction between 2-acylnaphthoquinones and 2-aminobenzothiazoles. A new synthetic route to antiproliferative benzo[g]benzothiazolo[2,3-b]quinazoline-7,12-quinones. Tetrahedron Lett. 56 , 5103–5105. 10.1016/j.tetlet.2015.07.034 Search in Google Scholar

Valenzuela, M., Glorieux, C., Stokis, J., Sid, B., Sandoval, M., Felipe, K.B., Kviecinski, M., Verrax, J., and Buc Calderon, P. (2014). Retinoic acid synergizes ATO-mediated cytotoxicity by precluding Nrf2 activity in AML cells. Br. J. Cancer 111 , 874–882. 10.1038/bjc.2014.380 Search in Google Scholar

van Heyningen, V., Boyd, P.A., Seawright, A., Fletcher, J.M., Fantes, J.A., Buckton, K.E., Spowart, G., Porteous, D.J., Hill, R.E., and Newton. M.S. (1985). Molecular analysis of chromosome 11 deletions in aniridia-Wilms tumor syndrome. Proc. Natl. Acad. Sci. USA 82 , 8592–8596. 10.1073/pnas.82.24.8592 Search in Google Scholar

Venkatesan, B., Mahimainathan, L., Das, F., Ghosh-Choudhury, N., and Ghosh, C.G. (2007). Downregulation of catalase by reactive oxygen species via PI 3 kinase/Akt signaling in mesangial cells. J. Cell Physiol. 211 , 457–467. 10.1002/jcp.20953 Search in Google Scholar

Verrax, J. and Buc Calderon, P. (2009). Pharmacologic concentrations of ascorbate are achieved by parenteral administration and exhibit antitumoral effects. Free Radic. Biol. Med. 47 , 32–40. 10.1016/j.freeradbiomed.2009.02.016 Search in Google Scholar

Verrax, J., Cadrobbi, J., Delvaux, M., Jamison, J.M., Gilloteaux, J., Summers, J.L., Taper, H.S., and Buc Calderon, P. (2003). The association of vitamins C and K3 kills cancer cells mainly by autoschizis, a novel form of cell death. Basis for their potential use as coadjuvants in anticancer therapy. Eur. J. Med. Chem. 38 , 451–457. 10.1016/S0223-5234(03)00082-5 Search in Google Scholar

Verrax, J., Cadrobbi, J., Marques, C., Taper, H., Habraken, Y., Piette, J., and Buc Calderon, P. (2004). Ascorbate potentiates the cytotoxicity of menadione leading to an oxidative stress that kills cancer cells by a non-apoptotic caspase-3 independent form of cell death. Apoptosis 9 , 223–233. 10.1023/B:APPT.0000018804.26026.1a Search in Google Scholar

Verrax, J., Delvaux, M., Beghein, N., Taper, H., Gallez, B., and Buc Calderon, P. (2005). Enhancement of quinone redox cycling by ascorbate induces a caspase-3 independent cell death in human leukaemia cells. An in vitro comparative study. Free Radic. Res. 39 , 649–657. 10.1080/10715760500097906 Search in Google Scholar

Verrax, J., Stockis, J., Tison, A., Taper, H.S., and Buc Calderon, P. (2006). Oxidative stress by ascorbate/menadione association kills K562 human chronic myelogenous leukaemia cells and inhibits its tumour growth in nude mice. Biochem. Pharmacol. 72 , 671–680. 10.1016/j.bcp.2006.05.025 Search in Google Scholar

Verrax, J., Stockis, J., Vanbever, S., Taper, H., and Buc Calderon, P. (2007). Role of glycolysis inhibition and Poly(ADP-ribose) polymerase activation in necrotic-like cell death caused by ascorbate/menadione-induced oxidative stress in K562 human chronic myelogenous leukemic cells. Int. J. Cancer 120 , 1192–1197. 10.1002/ijc.22439 Search in Google Scholar

Verrax, J., Pedrosa, R.C., Beck, R., Dejeans, N., Taper, H., and Buc Calderon, P. (2009). In situ modulation of oxidative stress: a novel and efficient strategy to kill cancer cells. Curr. Med. Chem. 16 , 1821–1830. 10.2174/092986709788186057 Search in Google Scholar

Verrax, J., Dejeans, N., Sid, B., Glorieux, C., and Buc Calderon, P. (2011a). Intracellular ATP levels determine cell death fate of cancer cells exposed to both standard and redox chemotherapeutic agents. Biochem. Pharmacol. 82 , 1540–1548. 10.1016/j.bcp.2011.07.102 Search in Google Scholar

Verrax, J., Beck, R., Dejeans, N., Glorieux, C., Sid, B., Pedrosa, R.C., Benites, J., Vasquez, D., Valderrama, J.A., and Buc Calderon, P. (2011b). Redox-active quinones and ascorbate: an innovative cancer therapy that exploits the vulnerability of cancer cells to oxidative stress. Anticancer Agents Med. Chem. 11 , 213–221. 10.2174/187152011795255902 Search in Google Scholar

Vetrano, A.M., Heck, D.E., Mariano, T.M., Mishin, V., Laskin, D.L., and Laskin, J.D. (2005). Characterization of the oxidase activity in mammalian catalase. J. Biol. Chem. 280 , 35372–35381. 10.1074/jbc.M503991200 Search in Google Scholar

Vlasits, J., Jakopitsch, C., Schwanninger, M., Holubar, P., and Obinger, C. (2007). Hydrogen peroxide oxidation by catalase-peroxidase follows a non-scrambling mechanism. FEBS Lett. 581 , 320–324. 10.1016/j.febslet.2006.12.037 Search in Google Scholar

Wanders, R.J., Kos, M., Roest, B., Meijer, A.J., Schrakamp, G., Heymans, H.S., Tegelaers, W.H., van den Bosch, H., Schutgens, R.B., and Tager, J.M. (1984). Activity of peroxisomal enzymes and intracellular distribution of catalase in Zellweger syndrome. Biochem. Biophys. Res. Commun. 123 , 1054–1061. 10.1016/S0006-291X(84)80240-5 Search in Google Scholar

Wang, C.D., Sun, Y., Chen, N., Huang, L., Huang, J.W., Zhu, M., Wang, T., and Ji, Y.L. (2016). The role of catalase C262T gene polymorphism in the susceptibility and survival of cancers. Sci. Rep. 6 , 26973. 10.1038/srep26973 Search in Google Scholar

Warburg, O. (1923). Versuche an uberlebendem karzinomgewebe. Biochem. Z. 142 , 317–333. Search in Google Scholar

Welsh, S.J., Williams, R.R., Birmingham, A., Newman, D.J., Kirkpatrick, D.L., and Powis, G. (2003). The thioredoxin redox inhibitors 1-methylpropyl 2-imidazolyl disulfide and pleurotin inhibit hypoxia-induced factor 1alpha and vascular endothelial growth factor formation. Mol. Cancer Ther. 2 , 235–243. Search in Google Scholar

Wieacker, P., Mueller, C.R., Mayerova, A., Grzeschik, K.H., and Ropers, H.H. (1980). Assignment of the gene coding for human catalase to the short arm of chromosome 11. Ann. Genet. 23 , 73–77. Search in Google Scholar

Wink, D.A. and Mitchell, J.B. (1998). Chemical biology of nitric oxide: insights into regulatory, cytotoxic, and cytoprotective mechanisms of nitric oxide. Free Radic. Biol. Med. 25 , 434–456. 10.1016/S0891-5849(98)00092-6 Search in Google Scholar

Winternitz, M.C. and Meloy, C.R. (1908). On the occurrence of catalase in human tissues and its variations in diseases. J. Exp. Med. 10 , 759–781. 10.1084/jem.10.6.759 Search in Google Scholar

Yano, S., Arroyo, N., and Yano, N. (2004a). Catalase binds Grb2 in tumor cells when stimulated with serum or ligands for integrin receptors. Free Radic. Biol. Med. 36 , 1542–1554. 10.1016/j.freeradbiomed.2004.04.006 Search in Google Scholar

Yano, S., Arroyo, N., and Yano, N. (2004b). SHP2 binds catalase and acquires a hydrogen peroxide-resistant phosphatase activity via integrin-signaling. FEBS Lett. 577 , 327–332. 10.1016/j.febslet.2004.10.011 Search in Google Scholar

Zamocky, M. and Koller, F. (1999). Understanding the structure and function of catalases: clues from molecular evolution and in vitro mutagenesis. Prog. Biophys. Mol. Biol. 72 , 19–66. 10.1016/S0079-6107(98)00058-3 Search in Google Scholar

Zamocky, M., Furtmuller, P.G., and Obinger, C. (2008). Evolution of catalases from bacteria to humans. Antioxid. Redox. Signal. 10 , 1527–1548. 10.1089/ars.2008.2046 Search in Google Scholar PubMed PubMed Central

Zamocky, M., Furtmuller, P.G., and Obinger, C. (2010). Evolution of structure and function of Class I peroxidases. Arch. Biochem. Biophys. 500 , 45–57. 10.1016/j.abb.2010.03.024 Search in Google Scholar PubMed

©2017 Walter de Gruyter GmbH, Berlin/Boston

  • X / Twitter

Supplementary Materials

Please login or register with De Gruyter to order this product.

Biological Chemistry

Journal and Issue

Articles in the same issue.

redox reaction and catalase activity experiment

Home

  • Terms of Use
  • Privacy Policy
  • Internet Cookies
  • Instructional Exchange
  • Educator's Toolkit
  • Ayers Exclusives
  • Tennessee Tools
  • Education Events
  • Conference Materials
  • Crosswalk Charts
  • Podcast Resources
  • Webinar Sessions

Enzyme Action: Catalase Lab

This lab experiment is posted on the Vernier website. Vernier is one of the leading producers of probeware and software for data collection in classrooms and labs. There are several student lab procedures you can download directly from this site, but to access the teacher background information you must purchase the lab book entitle Biology with Vernier . Obviously, this lab is only applicable if your school has access to the Vernier probes. If you have access to Vernier probes, this lab is an excellent way to emphasize this particular course level expectation. Hydrogen peroxide is lethal to cells in high doses, so our cells produce the enzyme catalase in order to break hydrogen peroxide down into water and oxygen. The probes in the lab will be measuring the oxygen output via pressure. The more pressure, the more oxygen produced. This lab has three different variables: 1) concentration of the enzyme; 2) temperature of the environment; and 3) pH of the environment. It is possible to complete one section per class period. The data collection is superb and lends itself well to a formal lab report.

Standards & Objectives

In this experiment, you will

  • Use a computer and Gas Pressure Sensor to measure the production of oxygen gas as hydrogen peroxide is destroyed by the enzyme catalase or peroxidase at various enzyme concentrations.
  • Measure and compare the initial rates of reaction for this enzyme when different concentrations of enzyme react with H2O2.
  • Measure the production of oxygen gas as hydrogen peroxide is destroyed by the enzyme catalase or peroxidase at various temperatures.
  • Measure and compare the initial rates of reaction for the enzyme at each temperature.
  • Measure the production of oxygen gas as hydrogen peroxide is destroyed by the enzyme catalase or peroxidase at various pH values.
  • Measure and compare the initial rates of reaction for the enzyme at each pH value.

Part I Effect of Enzyme Concentration

  • How does changing the concentration of enzyme affect the rate of decomposition of H2O2?
  • What do you think will happen to the rate of reaction if the concentration of enzyme is increased to five drops? Predict what the rate would be for 5 drops.

Part II Effect of Temperature

  • At what temperature is the rate of enzyme activity the highest? Lowest? Explain.
  • How does changing the temperature affect the rate of enzyme activity? Does this follow a pattern you anticipated?
  • Why might the enzyme activity decrease at very high temperatures?

Part III Effect of pH

  • At what pH is the rate of enzyme activity the highest? Lowest?
  • How does changing the pH affect the rate of enzyme activity? Does this follow a pattern you anticipated? 

Lesson Variations

  • Different organisms often live in very different habitats. Design a series of experiments to investigate how different types of organisms might affect the rate of enzyme activity.
  • Consider testing a plant, an animal, and a protist.
  • Presumably, at higher concentrations of H2O2, there is a greater chance that an enzyme molecule might collide with H2O2. If so, the concentration of H2O2 might alter the rate of oxygen production. Design a series of experiments to investigate how differing concentrations of the substrate hydrogen peroxide might affect the rate of enzyme activity.
  • Design an experiment to determine the effect of boiling the catalase on the reaction rate.
  • Explain how environmental factors affect the rate of enzyme-catalyzed reactions. 

Helpful Hints

  • computer 
  • 600 mL beaker
  • Vernier computer interface 
  • enzyme suspension
  • LoggerPro 
  • four 18 X 150 mm test tubes
  • Vernier Gas Pressure Sensor 
  • 1-hole rubber stopper assembly 
  • 10 mL graduated cylinder 
  • test tube rack
  • 250 mL beaker of water 
  • thermometer
  • 3% H2O2 
  • three dropper pipettes

redox reaction and catalase activity experiment

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
  • Oxid Med Cell Longev
  • v.2019; 2019

Logo of oximed

Role of Catalase in Oxidative Stress- and Age-Associated Degenerative Diseases

Ankita nandi.

1 Department of Biotechnology, Visva-Bharati University, Santiniketan, West Bengal 731235, India

Liang-Jun Yan

2 Department of Pharmaceutical Sciences, UNT System College of Pharmacy, University of North Texas Health Science Center, Fort Worth, TX 76107, USA

Chandan Kumar Jana

3 Department of Chemistry, Purash-Kanpur Haridas Nandi Mahavidyalaya, P.O. Kanpur, Howrah, West Bengal 711410, India

Nilanjana Das

Associated data.

Reactive species produced in the cell during normal cellular metabolism can chemically react with cellular biomolecules such as nucleic acids, proteins, and lipids, thereby causing their oxidative modifications leading to alterations in their compositions and potential damage to their cellular activities. Fortunately, cells have evolved several antioxidant defense mechanisms (as metabolites, vitamins, and enzymes) to neutralize or mitigate the harmful effect of reactive species and/or their byproducts. Any perturbation in the balance in the level of antioxidants and the reactive species results in a physiological condition called “oxidative stress.” A catalase is one of the crucial antioxidant enzymes that mitigates oxidative stress to a considerable extent by destroying cellular hydrogen peroxide to produce water and oxygen. Deficiency or malfunction of catalase is postulated to be related to the pathogenesis of many age-associated degenerative diseases like diabetes mellitus, hypertension, anemia, vitiligo, Alzheimer's disease, Parkinson's disease, bipolar disorder, cancer, and schizophrenia. Therefore, efforts are being undertaken in many laboratories to explore its use as a potential drug for the treatment of such diseases. This paper describes the direct and indirect involvement of deficiency and/or modification of catalase in the pathogenesis of some important diseases such as diabetes mellitus, Alzheimer's disease, Parkinson's disease, vitiligo, and acatalasemia. Details on the efforts exploring the potential treatment of these diseases using a catalase as a protein therapeutic agent have also been described.

1. Introduction

Reactive species (RS) are highly active moieties, some of which are direct oxidants, and some have oxygen or oxygen-like electronegative elements produced within the cell during cellular metabolism or under pathological conditions. Some of the reactive species are free radicals such as the hydroxyl radical and the superoxide radical, and some are nonradicals such as hydrogen peroxide. Free radicals are any independent species which consist of one or more unpaired electrons in their atomic or molecular orbital. They are generally unstable, short lived, but usually chemically reactive. They can react with any molecule either by oxidizing it or by causing any other kind of chemical modification. Free radicals can potentially oxidize all cellular biomolecules including nucleic acids, proteins, and lipids. For example, peroxidation of omega-6 polyunsaturated fatty acid (such as arachidonic acid and linoleic acid) leads to the production of 4-hydroxynonenal (HNE), which is one of the main reactive aldehydes produced by oxidative stress [ 1 ]. There are many reactive species and free radicals [ 2 ] which are listed in Table 1 .

Examples of the various free radicals and other oxidants in the cell [ 2 ].

Reactive oxidantsExamples
Reactive oxygen species (ROS)Superoxide (O ), hydroxyl radical (OH ), hydrogen peroxide (H O ), alkoxyl radical (RO ), lipid alkoxyl (LO ), Peroxyl radical (RO2 ), ozone (O ), lipid peroxide (LOOH), singlet oxygen ( O ), Hydroperoxyl radical (HO )
Reactive chlorine species (RCS)Hypochlorite ion (OCl ), nitryl chloride (NO Cl)
Reactive nitrogen species (RNS)Nitric oxide (NO ), nitrous acid (HNO ), Nitrosonium cation (NO ), nitrosyl anion (NO ), peroxynitrite (ONOO ), nitrogen dioxide (NO ), alkyl peroxynitrite (ROONO)
Reactive sulfur species (RSS)Thiyl radical (R-S ), perthiyl radical (RSS )

These free radicals are formed in the cell during normal cellular metabolism as mitochondrial electron transport chain, β -oxidation of fatty acids, and cytochrome P450-mediated reactions and by the respiratory burst during immune defense. For example, autooxidation of some biologically important substances such as FADH 2 and tetrahydropteridines can yield O 2 · – in the presence of oxygen [ 3 ]. The imbalance between production and quenching of these reactive substances through antioxidant mechanisms causes oxidative stress. The loss of functionality and adaptability of important biomolecules due to oxidative stress are two interdependent biological processes, which are among the important factors that mediate aging. The free radical hypothesis, also known as oxidative stress hypothesis, is one of the strongly supported theories which can define the causes behind the aging process.

Oxidative stress has been implicated in many metabolic and neurologic degenerative disorders. Degenerative diseases, where the function and structure of a tissue or organs deteriorate over time such as in Alzheimer's disease, Parkinson's disease, diabetes, cataracts, cancer, and cardiovascular disease, have been attributed to oxidative stress conditions and the process of natural aging. Thus, oxidative stress, aging, and degenerative diseases are interconnected.

The body has a defense mechanism against oxidative stress in which both enzymatic and nonenzymatic molecules are the two prime components. This antioxidant defense system consists of some enzymes, some proteins, and a few low molecular weight molecules. The antioxidant enzymes can catalytically remove the reactive species. For example, superoxide dismutase dismutates superoxide into hydrogen peroxide which is in turn degraded by catalase or by glutathione peroxidase. The relationship between the different antioxidant enzymes is depicted graphically in Figure 1 . Transferrin, metallothionein, and caeruloplasmin are some of the proteins which can reduce the availability of prooxidants such as transition metal ions like iron ions and copper ions which can produce a hydroxyl radical from hydrogen peroxide by the Fenton reaction. The low molecular weight antioxidants include ascorbic acid, α -tocopherol, glutathione, and uric acid, which neutralize the RS by scavenging the whole molecule or its byproducts, by reducing it by or participating in any form of chemical reaction leading to complete or partial destruction of it or its byproducts. The interaction of catalase with other antioxidants and proteins can be predicted by the STRING (Search Tool for the Retrieval of Interacting Genes/Proteins) analysis [ 4 , 5 ]. STRING is a biological database used to study protein-protein interaction. The STRING network analysis of catalase's interaction with other proteins has been categorized into two distinct modules ( Figure 2 ). Module 1 contains four proteins which are basically involved in the pathways of peroxisomes including CAT and three proteins of module 2 such as SOD1 (superoxide dismutase 1), SOD2 (superoxide dismutase 2), and PRDX5 (peroxiredoxin 5) [ 6 – 8 ] (Supplementary ). In module 1, ACOX1 (peroxisomal acyl coenzyme A oxidase), HSD17B4 (peroxisomal multifunctional enzyme), and HAO1 (hydroxyacid oxidase 1) are involved in the fatty acid oxidation pathway in the peroxisome while the protein DAO (D amino acid oxidase) is involved in the amino acid metabolism pathway in the peroxisome [ 6 – 8 ] (Supplementary ). All the components of module 1 are involved in different metabolic pathways. The proteins in module 2 are mainly involved in responses against oxidative stress. All the proteins have antioxidant activity except AKT1 (RAC-alpha serine-threonine protein kinase). AKT1 is a serine-threonine protein kinase which is involved in cell survival, metabolism, growth, and angiogenesis. All the proteins of both modules 1 and 2 including CAT have catalytic activity and are located in the lumen of intracellular organelles. SOD2 and AKT1 of module 2 including CAT were involved in the longevity regulating pathway and FOXO signaling pathway in mammals [ 6 – 8 ] (Supplementary Figures and ). But in multiple other species, SOD1 and SOD3 (superoxide dismutase 3) were also involved along with SOD2, AKT1, and CAT [ 6 – 8 ] (Supplementary ). Among the reactive species, hydrogen peroxide is freely diffusible and is relatively long-lived. It acts as a weak oxidizing as well as reducing agent; however, it is not very reactive, but it is the progenitor of many other reactive oxygen species (ROS). It has been demonstrated to oxidatively modify glyceraldehyde-3-phosphate dehydrogenase by oxidation of the labile essential thiol groups at the active site of this enzyme [ 2 ]. In most cellular injuries, this molecule is known to play an indirect role. One of the most important products is the formation of a more reactive free radical · OH radical in the presence of transition metal ions such as Fe 2+ by means of the Fenton reaction.

An external file that holds a picture, illustration, etc.
Object name is OMCL2019-9613090.001.jpg

Relationship between catalase and other antioxidant enzymes.

An external file that holds a picture, illustration, etc.
Object name is OMCL2019-9613090.002.jpg

STRING analysis of interaction of catalase with other proteins.

There are many enzymes that are able to neutralize hydrogen peroxide. These enzymes include catalase, glutathione peroxidase, and other peroxidases such as cytochrome c peroxidase and NADH peroxidase [ 2 ]. Catalase is a key enzyme which uses hydrogen peroxide, a nonradical ROS, as its substrate. This enzyme is responsible for neutralization through decomposition of hydrogen peroxide, thereby maintaining an optimum level of the molecule in the cell which is also essential for cellular signaling processes. The importance of the enzyme can be gauged from the fact of its direct and indirect involvement in many diseases and infections. In this review, an attempt has been made to correlate the role of catalase with the pathogenesis and progression of oxidative stress-related diseases. A brief account of catalase, its isoforms, structure, and reaction mechanism, and its relation with some common important disorders is described in this review article.

2. Catalase

A catalase (E.C. 1.11.1.6) is one of the most important antioxidant enzymes. It is present in almost all aerobic organisms. Catalase breaks down two hydrogen peroxide molecules into one molecule of oxygen [ 9 ] and two molecules of water in a two-step reaction [ 10 ]. The same is represented in Figure 3 as derived from Ivancich et al. [ 11 ] and Lardinois [ 12 ]. The first step of the reaction mechanism involves formation of a spectroscopically distinct intermediate compound I ( Figure 3(a) ) which is a covalent oxyferryl species (Fe IV O) having a porphyrin π -cation radical, through the reduction of one hydrogen peroxide molecule [ 11 ]. In the second step reaction ( Figure 3(b) ), compound I is reduced through redox reactions by a two-electron transfer from an electron donor (the second molecule of hydrogen peroxide) to produce the free enzyme, oxygen, and water [ 10 ].

An external file that holds a picture, illustration, etc.
Object name is OMCL2019-9613090.003.jpg

Steps in catalase reaction: (a) first step; (b) second step.

In 1937, the protein was first crystallized from bovine liver at Sumner and Dounce's laboratory [ 13 ]. The first prokaryotic catalase was purified from an aerobic bacterium, Micrococcus lysodeikticus , in 1948 [ 14 ]. The gene coding for catalase is the CAT gene which is positioned in chromosome 11 in humans. In the following decades, several studies have been carried out on prokaryotic catalase and also on the lower eukaryotic catalase. In particular, research on catalase from Saccharomyces cerevisiae has generated data and information on the evolution of the enzyme at the molecular level. It has also been reported that catalase is an important enzyme implicated in mutagenesis and inflammation conditions as well as during the suppression of apoptosis [ 15 – 18 ] which are all known to be associated with oxidative stress conditions.

Catalase has been characterized from many eukaryotic as well as prokaryotic organisms. Table 2 summarizes some basic physiochemical information available in the literature to date on catalase from different organisms. Based on the differences in their sequence and structure, there are three different types of catalase. The monofunctional heme-containing enzyme is the most widespread one. It is present in all aerobic organisms. The bifunctional catalase-peroxidase belongs to the second class, which is relatively less abundant in nature. This enzyme also contains a heme group. It is closely related to the plant peroxidases with structural and sequence similarities. The third class belongs to the Mn-containing catalase group which lacks the heme group.

Physicochemical characteristics of catalase from various sources.

Organisms/organ/organelleSpecific activity ( mol/min/mg)Optimum temperature (°C)/pHInhibitors value (mM)Turnover numberMol. wt.Reference
erythrocyte (cytoplasm), kidney, and liver (mitochondria, peroxisome)27380037°C/6.8-7.53-Amino-1-H-1,2,4-triazole80--[ , ]
liver9180025°C-35°C/6-7.53-Amino-1-H-1,2,4-triazole28.6--[ – ]
-25°C/6-10Hydrogen peroxide (above 60 mM)10080000234000[ , ]
seedling2570040°C/7Cu , Fe , EDTA, NaN 16.2-[ ]
2070022°C/6-82-Mercaptoethanol6416300337000[ , ]
116100-NaCN (35 mM), hydroxylamine125-[ ]

Humans possess a typical monofunctional heme-containing catalase having a prosthetic group of ferric protoporphyrin IX which reacts with hydrogen peroxide. Located in the peroxisomes, the enzyme has a molecular mass of approximately 220-240 kDa [ 19 ]. It is a tetrameric protein with each subunit divided into four domains, the N-terminal threading arm, C-terminal helices, wrapping loop, and β barrel [ 20 ] ( Figure 4 ). Each subunit has a hydrophobic core comprising eight stranded β barrels surrounded by α -helices. These β barrels are antiparallel with each other. The heme distal side of the subunit is made up of the first four β strands ( β 1- β 4) of the β barrel domain and the remaining four strands ( β 5- β 8) play a part in the NADPH binding pocket. The N-terminal threading arm of a subunit (residues 5-70) intricately connects two subunits by hooking through a long wrapping loop (residues 380-438). Finally, a helical domain at one face of the β barrel is composed of four C-terminal helices. Tetramerization forces the N-terminal threading arms from the arm-exchanged dimer to cover the heme active site for the other pair of dimers and suggests that catalase fits the more general pattern of domain swapping with the arm-exchange being a later, tetramer-dependent elaboration. Throughout the protein, water fills in packing defects between the four domains of the subunit and between subunits within the tetramer. Only the hydrophobic β barrel and the immediate vicinity of the active site are substantially devoid of these structural water molecules. From XRD studies, the root mean square deviation (r.m.s.d.) coordinate differences between the four subunits were found to be 0.156 Å for the backbone atoms, 0.400 Å for the side chains, and 0.125 Å for the heme groups [ 21 ]. The 3D structure of the enzyme at 1.5 Å was elucidated in 2001 [ 22 ]. The crystal structures of human catalase show that the active site iron is pentacoordinated. The negatively charged heme carboxylate radical forms salt bridges to three arginine residues (Arg72, Arg117, and Arg365) which likely aid in heme burial and help increase the redox potential of the compound I porphyrin radical and are conserved in bacterial, fungal, plant, and animal catalase. Besides the heme group, the active conformation of the enzyme consists of one tightly bound NADPH molecule in each subunit. There are various reports on the role of this NADPH molecule. It has been demonstrated to obstruct the formation of Fe (IV)oxo-ligated porphyrin, an inactive form of catalase—by hydrogen peroxide, and to also slowly induce the removal of inactive catalase [ 19 , 23 ].

An external file that holds a picture, illustration, etc.
Object name is OMCL2019-9613090.004.jpg

Crystal structure of human erythrocytic catalase [ 20 ] PDB ID: 1F4J.

3. Catalase-Related Diseases

Catalase deficiency or malfunctioning is associated with many diseases such as diabetes mellitus, vitiligo, cardiovascular diseases, Wilson disease, hypertension, anemia, some dermatological disorders, Alzheimer's disease, bipolar disorder, and schizophrenia [ 24 – 26 ] ( Figure 5 ). It has been reported that an anomaly of catalase activity is inherited in acatalasemia which is a rare genetic disorder (also known as Takahara disease) [ 27 ]. It is an autosomal recessive trait and characterized by a reduced level of catalase. Catalase has a prime role in regulating the cellular level of hydrogen peroxide [ 28 , 29 ], and its hydrogen peroxide catabolism protects the cells from oxidative assault, for example, by securing the pancreatic β cells from hydrogen peroxide injury [ 30 , 31 ]. Low catalase activities have been reported in schizophrenic patients such as also in patients with atherosclerosis [ 32 ].

An external file that holds a picture, illustration, etc.
Object name is OMCL2019-9613090.005.jpg

List of some diseases linked to catalase deficiency.

Genetic variations in the catalase gene and in its promoter region also play a role in the pathogenesis of various diseases which is depicted in Figure 6 . Several studies have investigated CAT polymorphism and its involvement in the development of various diseases as well as its role as regulator in the CAT gene expression. Single nucleotide polymorphisms of the CAT gene in the promoter region possibly affect the transcription frequencies resulting in low CAT expression [ 33 , 34 ]. The most common polymorphisms that influence the transcription of the CAT gene and also affect catalase activity are -262C/T and -844G/A or -844C/T in the promoter region [ 35 ]. There are many other polymorphisms involved in the development of numerous diseases which varies amongst populations. CAT -262C/T polymorphism is related to type 1 diabetes and breast cancer [ 36 – 38 ]. Two single nucleotide polymorphisms of CAT gene, viz., 1167T/C and -262C/T, have been reported to have a strong association with type 1 diabetes mellitus [ 36 ]. The functional consequence of this 1167T/C polymorphic positioned in exon 9 is not known. But in case of -262C/T, the variation shows significant functional significance. It influences the AP-2 and Sp-1 (nuclear transcriptional factors) binding and also effects the expression as well as the level of catalase in the red blood cells [ 36 ]. In Swedish populations, the concentration of erythrocytic catalase in individuals carrying the TT genotype was high compared to those of the CC genotype [ 39 ]. In Russian populations—on the other hand—the individuals carrying the CC genotype have a higher risk of developing type 1 diabetes than those carrying the TT genotype [ 36 ]. The blood catalase level was found to be low in CC individuals which results in oxidative stress conditions, thereby promoting type 1 diabetes [ 36 ]. Another single nucleotide polymorphism of the CAT gene 111C/T in exon 9 was examined among different forms of diabetes and showed a very poor association [ 40 ]. The CAT -262C/T polymorphism has an association with breast cancer. The CC genotype showed higher catalase activity in red blood cell as compared to TT and TC genotypes with a correlated reduced risk of breast cancer by 17% [ 38 ]. However, it must be noted here that this population study was performed with a much lesser number of individuals. Studies have shown that the level of -262C/T polymorphism effects not only the transcriptional activity but also the level of catalase in red blood cells [ 37 ]. Another common CAT polymorphism is -844C/T or -844G/A which might result in a lower catalase level by influencing the transcription frequency. CAT -844C/T polymorphism has a strong association with hypertension among the Chinese population [ 41 ]. Hypertension is a multifactorial complex lifestyle disease. Among Japanese populations, this -844C/T polymorphism has been reported to show a strong association with hypertension [ 42 ]. But the functional relationship is not very clear. The CAT -844G/A, -89A/T, and -20T/C polymorphisms have been shown to be associated with malnutrition [ 43 ]. This polymorphism might affect the transcription rates thereby lowering the catalase level. The -89A/T polymorphism has also been reported to exhibit an association with vitiligo and osteonecrosis [ 44 , 45 ]. The variant with CAT -89A/T has been reported to be associated with a significantly reduced level of catalase with a correlation with developing vitiligo in the Chinese population [ 44 ]. The CAT 389C/T genotype has no reported association with vitiligo in the Chinese population, but a connection has been established in North America and the United Kingdom [ 44 , 46 , 47 ]. The relation of these genotypes with vitiligo pathogenesis is discussed in a later section. The CAT -89A/T, -20T/C, +3033C/T, +14539A/T, +22348C/T, and +24413T/C polymorphisms might be involved in osteonecrosis amongst Korean populations [ 45 ]. Data from all the studies show different polymorphisms of the CAT gene among different populations in various regions of the world. Further population-based research across the world is required to gain a clear idea about the association of the CAT gene in different diseases. In the future, this might unlock new therapeutic approaches by regulating the CAT gene.

An external file that holds a picture, illustration, etc.
Object name is OMCL2019-9613090.006.jpg

Association of catalase polymorphism with risk of some widespread diseases.

3.1. Diabetes Mellitus

Diabetes mellitus is a common disease nowadays. They are caused by a bundle of metabolic disorders, distinguished by high levels of glucose in the blood due to improper secretion of insulin or its activity or both. It can lead to other secondary afflictions such as nerve damage, blindness, heart disease, stroke, and kidney disease. There has been a significant rise in the diabetes-affected population in recent years. It is estimated that, worldwide, the number of diabetes-affected adults will increase more than twofold from the 135 million affected in 1995 to approximately 300 million by 2025 [ 48 ] and 629 million by 2045 [ 49 ], and the majority of increment will be from developing countries such as India [ 48 ].

The 2018 data from the diabetes country profile from the World Health Organization (WHO) [ 50 ] is depicted in Figure 7 which shows the prevalence of diabetes amongst both genders in different countries classified according to their economic status by a United Nation's report. The disease seems more prevalent in the developed nations, and the percentage of the affected population seems to show a more or less uniform level in all these countries. A lot of discrepancies in the level are observed amongst the developing nations with the highest percentage of the population being affected in Egypt. Less prevalence is observed amongst the least developing nations indicating that lifestyle and diet play a major role in development of the disease as observed amongst the developed nations. Gender also seems to play a role with more prevalence of the disease among the females than the males in the developing nations indicating that societal norms may also play a role. In contrast, males seem more susceptible in developed nations, which indicates a possible genetic and lifestyle role in development of the disease.

An external file that holds a picture, illustration, etc.
Object name is OMCL2019-9613090.007.jpg

Prevalence of diabetes amongst males and females in some countries in 2018 (data source: World Health Organization-Diabetes Country Profile 2018).

There are two general forms of diabetes mellitus, type 1 and type 2. Type 1 diabetes mellitus is a juvenile form and insulin-dependent diabetes which accounts for approximately 10% of all cases, but it may also develop in adults [ 51 ]. In this case, pancreatic β cells are destroyed by autoantibodies rendering the cells incapable of producing insulin. This autoimmune disease has a correlation between immunologic and genetic factors. There are three major types of autoantibodies found in type 1 diabetes such as GDP65, IA2, and insulin autoantibodies, but antibodies against insulin can be identified mostly in young patients and may be lacking in adults [ 52 , 53 ]. These antibodies bind mainly to the conformational epitopes on the B chain of insulin. The genetic feature shows a relationship between type 1 diabetes and some alleles of the HLA complex. There is a strong connection between the progression of type 1 diabetes and the presence of HLA class II alleles.

Type 2 diabetes mellitus is the most common form of the disease, accounting for approximately 90% of all diabetes cases. It occurs primarily due to low production of insulin and secondarily also due to insulin resistance by the body's cells. The β cells of islets of Langerhans become damaged which make them unable to produce insulin. Oxidative stress has been demonstrated to be an important factor responsible for the advancement of type 2 diabetes. It has been demonstrated that hydrogen peroxide acts as an oxidant and damages the β cell interrupting the signaling pathway of insulin production [ 30 , 54 , 55 ]. According to a study from Prof. Kassab's laboratory, a four-fold increase in the concentration of hydrogen peroxide was observed in type 2 diabetes mellitus patients than in the healthy controls [ 56 ]. This observation was corroborated with observations of low catalase activity in the β cells in hyperglycemic mice models [ 57 ].

Another form of diabetes known as pancreatogenic diabetes has been classified as type 3c diabetes mellitus (T3cDM). T3cDM is the result of pancreatitis (both acute and chronic), cystic fibrosis in the tissue of pancreas, inflammation, and damage of pancreatic tissue [ 58 , 59 ]. The damage of exocrine pancreatic peptide (PP) and pancreatic enzymes occurs at the early phase of pancreatic diabetes. The reduction of the glycogen level due to damage of α cells occurs at a late phase of pancreatic diabetes. The resultant elevated level of glucagon can lead to hyperglycemia in diabetes mellitus [ 60 ]. There are many aspects associated with the pathophysiology of pancreatic diabetes. Immunopathogenesis is one of the important aspects which contribute to the development of pancreatic diabetes. Different proinflammatory cytokines like tumor necrosis factor α , interferon γ , and interleukin 1 β are involved in the pathogenesis of pancreatic diabetes [ 60 ]. Higher concentration of cytokines leads to the dysfunction of the β cells at an early stage of chronic pancreatitis [ 61 ]. At higher concentration, interleukin 1 β induces the apoptosis of β cells by the NF κ B pathway [ 62 ]. Higher concentration of interferon γ diminishes the translocation of pancreatic and duodenal home box 1 (PDX1), a transcription factor. PDX1 is important for the development of pancreatic cells through maturation of β cells and also via duodenal differentiation [ 63 ]. Reduction of survivability and differentiation of β cells occur in patients with chronic pancreatitis due to loss of PDX1. Hydrogen peroxide plays a central role in this pathway as a signaling molecule [ 64 ]. At lower concentration, hydrogen peroxide plays as a signaling molecule while it becomes toxic at higher concentration [ 65 ] and catalase plays an important role in maintaining homeostasis of the cells by degrading hydrogen peroxide. The activity of catalase in the serum was observed to be high in acute pancreatitis [ 66 ] and persists at its elevated level for as long as 10 to 14 days [ 66 ]. Therefore, the high catalase activity may contribute to the pathogenesis of T3cDM in an indirect way by maintaining the hydrogen peroxide concentration which would induce the synthesis of proinflammatory cytokines resulting in pancreatic diabetes.

Gestational diabetes mellitus (GDM) is another common form of diabetes among pregnant women. The pathogenesis of GDM is very similar to type 2 diabetes mellitus. There are several factors including ethnicity, maternal age, hypertension, obesity, and polycystic ovary syndrome (PCOS) which are associated with the possibility of developing GDM [ 67 , 68 ]. Pregnant women with GDM have higher risk of developing type 2 diabetes mellitus after pregnancy [ 68 ]. The offspring of gestational diabetic mothers are prone to development of different diseases like hypertension, different metabolic syndrome, and chronic kidney disease [ 69 , 70 ]. These birth defects might be due to higher concentration of reactive oxygen species and lowering of the antioxidant defense which in turn make the cell more susceptible to oxidative insults [ 70 , 71 ]. GDM usually develops in the second and third trimesters of the pregnancy period. Reports on the link of catalase with GDM are very conflicting. It has been reported that oxidative stress is high in the second and third trimesters of pregnancy and the catalase activity was also low during this period [ 72 , 73 ]. The blood catalase activity has been reported to be low in pregnant women with GDM compared to nonpregnant and pregnant nondiabetic healthy control women [ 72 ]. However, the blood catalase activity was observed to increase in the third trimester than in the second trimester in pregnant individuals with GDM [ 72 ]. In another study, low blood catalase activity has been observed in pregnant women with GDM [ 40 ]. As already mentioned, there is poor association between 111C/T polymorphism and different forms of diabetes mellitus which include GDM [ 40 ]. The mRNA expression of the CAT gene in the placenta of gestational diabetic pregnant women was found to be higher in comparison to that in normal pregnant women [ 74 ]. So it may be concluded from the above that catalase might have a relation with GDM pathophysiology during pregnancy, but further research to establish the facts is needed.

Hydrogen peroxide has been implicated to act as a cellular messenger in the signaling pathway for insulin secretion by inactivating tyrosine phosphatase [ 65 , 75 – 78 ]. It has been postulated that catalase in the liver may confer cellular protection by degrading the hydrogen peroxide to water and oxygen [ 28 – 31 ]. Lack of catalase can contribute to the development of diabetes mellitus [ 76 , 79 ] with a positive correlation being observed between diabetes mellitus in acatalasemic patients. It is estimated that approximately 12.7% of acatalasemic/hypocatalesemic patients are also affected by diabetes mellitus [ 79 ]. It was proposed that catalase deficiency may be responsible for the development of diabetes mellitus in an indirect way [ 24 ]. The β cells are known to be oxidant sensitive. These cells are not only deprived of catalase but also have a higher concentration of mitochondria [ 80 ] which is one of the major sources of superoxide and hydrogen peroxide in the cell through the electron transport pathway. Therefore, in acatalasemic/hypocatalesemic patients, a low amount of oxidative stress over a long period of time may result in the accumulation of oxidative damage in the β cells that results in the onset of diabetes [ 76 , 79 ].

There are many vascular complications in diabetes mellitus including microvascular complications (diabetic retinopathy, nephropathy, neuropathy, etc.) and cardiovascular complications [ 81 ]. Oxidation plays an important role in different complications which occur in both type 1 and type II diabetes. Due to the low expression levels or activity of catalase, the concentration of hydrogen peroxide may increase in the cells creating oxidative stress conditions causing the progression of different types of complications. In the case of diabetes retinopathy, the retina is damaged by retina neovascularization where new vessel origination from existing veins extends to the retinal inner cells [ 82 ] leading to blindness [ 83 ]. Vascular endothelial growth factor (VEGF) is a prime inducer of angiogenesis, a procedure of new vessel development. Nox4, a major isoform of NADPH oxidase, is predominant in the endothelial cells of the retina. It causes the generation of hydrogen peroxide instead of other reactive species [ 84 ]. Hydrogen peroxide may have a role as a signaling molecule in the VEGF signaling molecule. An upregulation of the Nox4 expression with downregulation of the catalase expression and/or activity in diabetes increases the hydrogen peroxide concentration which promotes retinal neovascularization through the VEGF signaling pathway [ 82 ]. In a study on a diabetic rat model, high concentration of hydrogen peroxide was observed in the retinal cells, creating oxidative stress conditions within the cell [ 85 ]. Since retinal cells have high content of polyunsaturated fatty acid content [ 86 ], they can be oxidized by the hydroxyl radicals generated from hydrogen peroxide by the Fenton reaction. High levels of lipid peroxides and oxidative DNA damage have been observed in diabetic retinopathy [ 87 – 90 ].

In a recent study, researchers have been able to distinguish five distinct clusters of diabetes by combining parameters such as insulin resistance, insulin secretion, and blood sugar level measurements with age of onset of illness [ 91 ]. Group 1 essentially corresponds to type 1 diabetes while type 2 diabetes is further subdivided into four subgroups labelled as group 2 to group 5. Individuals with impaired insulin secretion and moderate insulin resistance are labelled under group 2 (the severe insulin-deficient diabetes group) while in group 3, the severe insulin-resistant diabetes patients with obesity and severe insulin resistance are included. Group 4 is composed of the mild obesity-related diabetes patients who are obese and fall ill at a relatively young age while the largest group of patients is in group 5 with mild age-related diabetes in mostly elderly patients. A relationship between this new classification of diabetes with catalase expression levels or its activity has still not been probed for a link, if any, and needs further research.

3.2. Neurological Disorders

3.2.1. alzheimer's disease.

Alzheimer's disease is one of the onset of dementia diseases in adults [ 92 ]. According to the report of the Alzheimer's Association, approximately 5.5 million people in the United States of America were suffering from Alzheimer's disease in 2017. It is estimated that by 2050, the prevalence of Alzheimer's diseases will increase immensely from 4.7 million in 2010 to an estimated 13.8 million in 2050 [ 93 ].

Many factors including smoking and diabetes are associated with a higher risk of dementia. Alzheimer's disease is characterized by deposition of senile plaques of amyloid β peptides in the brain [ 94 , 95 ]. There are several studies which demonstrate that amyloid β peptides are toxic to neurons in culture [ 96 – 109 ]. Amyloid β , an amyloid precursor protein processing (APP) product, is a soluble component of the plasma and cerebrospinal fluid (CSF). In all cases of Alzheimer's disease, it has been observed that the soluble amyloid β is converted to insoluble fibrils in senile plaques through formation of protein-protein adducts [ 96 – 99 , 101 ].

It has been observed using in vitro cell culture studies that the nascent amyloid β is nontoxic but aged amyloid β becomes toxic to neurons [ 110 ]. It has been observed that amyloid β peptide is responsible for hydrogen peroxide accumulation within the cultures of neuroblastoma and hippocampal neurons [ 111 , 112 ] probably by the direct binding of amyloid β to catalase leading to decreased enzyme activities [ 26 ]. These findings led to development of the hypothesis that the catalase-amyloid β interaction may play a significant role in the increment of hydrogen peroxide in the cells linking the accretion of amyloid β and development of oxidative stress conditions in Alzheimer's disease [ 26 ]. So the current hypothesis regarding the mechanism of amyloid β -stimulated oxidative damage in cells is that amyloid β directly interacts with catalase by binding with the protein and deactivating its catalytic activity thereby creating oxidative stress conditions. In addition, full-length amyloid β peptides bind to Cu 2+ at their N-terminal section of the peptide and reduce it to the Cu + form [ 113 ]. It has been reported that amyloid β -Cu + complex can lead to hydrogen peroxide production [ 114 , 115 ]. Therefore, catalase has both a direct and an indirect relationship with the pathogenesis of Alzheimer's disease.

3.2.2. Parkinson's Disease

Parkinson's disease is an age-associated neurological disorder with the initial symptom as a simple tremor of the hand which gradually affects the whole body movement diminishing the quality of life severely with the advancement of the disease. Its clinical manifestations include bradykinesia, rigidity, resting tremor, and postural instability. It starts with rhythmic tremor of limbs especially during periods of rest or sleep. At the developing stage of the disease, patients face difficulties in controlling movement and muscle rigidity. Due to this muscular rigidity, slowness of movement and slowness of initiation of movement occur.

The disease is characterized by the exhaustion of dopamine due to damage of dopamine-producing neurons in the substantia nigra pars compacta (SNpc) [ 116 – 118 ]. It has been demonstrated that Parkinson's disease-affected patients suffer from 100-200 SNpc neuronal damages per day [ 119 ]. As various factors such as genetic inheritance, environmental toxins, oxidative stress, and mitochondrial dysfunction are probably involved in the pathogenesis of the disease, it is very challenging to understand the pathogenesis of Parkinson's disease.

It has been demonstrated that a protein, alpha ( α ) synuclein, is closely related to the cytopathology and histopathology of Parkinson's disease [ 120 ]. It has been observed that mutation in a gene responsible for the production of α -synuclein results in the production of a mutant protein that can promote the deposition of dopamine in the cytoplasm of neurons [ 121 ]. The small neurotransmitter molecules like dopamine are synthesized in the cytoplasm and are transferred to small vesicles as it becomes oxidized at the physiological pH. Mutant α -synuclein permeabilizes these vesicles causing leakage of the dopamine into the cytoplasm where it autooxidizes producing hydrogen peroxide, superoxide molecules, and toxic dopamine-quinone species creating oxidative stress conditions [ 122 ]. Mutant α -synuclein protein is also known to inhibit the expression and activity of catalase [ 123 ]. Arrest of catalase activity by α -synuclein is probably by hindering the peroxisome proliferator-activated receptor γ (PPAR γ ) transcription activity, which regulates the CAT gene expression [ 123 ]. Based on such experiments, it may be concluded that the low catalase activity and high hydrogen peroxide production in Parkinson's disease might be due to (the indirect) inhibition of catalase expression by the α -synuclein molecule.

3.3. Vitiligo

Vitiligo is one of the chronic pigmentary disorders where skin melanocyte cells—the pigment responsible for the color of the skin—are damaged or are unable to produce melanin. Various studies have shown that the catalase levels in the epidermis of vitiligo patients are lower as compared to those of the healthy control subjects [ 124 , 125 ] with a resultant increase in the concentration of hydrogen peroxide. In the cell, hydroxyl radicals can be produced spontaneously from hydrogen peroxide through photochemical reduction, i.e., the Haber-Weiss reaction [ 15 ]. These hydroxyl radicals are able to oxidize lipids in the cell membrane. This may be the cause behind damage of keratinocytes and melanocytes in the epidermal layer of the skin in such patients [ 126 – 130 ]. Moreover, the inhibitory effect of hydrogen peroxide or allelic modification of the CAT gene results in low catalase activity. However, it has been observed that there is an erratic relationship between catalase polymorphism and vitiligo. The 389C/T polymorphisms of exon 9, codon 389, and -89A/T of the promoter region were studied in vitiligo patients [ 34 , 44 , 46 , 47 , 131 , 132 ]. But the results were not observed to be consistent. Amongst the Chinese population, an association was observed in AT and TT genotypes with the increased risk of vitiligo whereas no association was observed between vitiligo and the -89A/T CAT polymorphism in the Korean population [ 34 , 44 ]. In the case of 389C/T polymorphism, several studies showed no difference between the controls and vitiligo patients [ 34 , 44 , 46 , 131 ] although contrary results have also been obtained in a few studies [ 47 , 131 ]. It has been reported that a mutation in the CAT gene might change the gene expression and/or cause structural changes in the keratinocytes and/or melanocytes [ 46 ]. Though the results are inconsistent from population studies, an interconnection between the pathogenesis and catalase may still be possible as scattered demonstrations are reported in the literature. Therefore, further studies to understand the link is necessary.

3.4. Acatalasemia

Acatalasemia (AC) is a hereditary disorder which is linked with the anomaly of catalase enzyme affecting its activity. In 1948, Takahara, a Japanese otolaryngologist, first reported this disorder [ 133 , 134 ]. He found that four out of seven races in Japan had the same genetic flaw [ 135 ]. His ex vivo experiments consisted of filling the mouth ulcer of a diseased patient with hydrogen peroxide. Since no bubble formation was observed, he concluded that a catalase or its enzymatic activity is absent in the saliva of the patients. In honor of his primary findings, this disease was christened as the Takahara disease. Acatalasemia and hypocatalasemia signify homozygotes and heterozygotes, respectively. The heterozygote of acatalasemia shows half of the catalase activity than normal and this phenotype is known as hypocatalasemia [ 136 ]. Depending on the geographical location from where it has been first studied, there are different types of acatalasemia described as Japanese, Swiss, Hungarian, German, and Peruvian types. Approximately 113 acatalasemic patients have been reported to date from all over the world.

Two kinds of mutations in the catalase gene have been reported to be involved in the Japanese acatalasemia. A splicing mutation has been held responsible for Japanese acatalasemia I where a substitution of a guanine residue with adenine residue at position 5 of intron 4 disturbed the splicing pattern of the RNA product producing a defective protein [ 137 ]. In Japanese acatalasemia II, a frame shift mutation occurs due to the deletion of thymine in position 358 of exon 4 which modifies the amino acid sequence and produces a new TGA (stop) codon at the 3′ terminal. Translation of this mutated strand produces a polypeptide of 133 amino acid residues. This is a truncated protein that is unstable and nonfunctional [ 138 ].

Aebi et al. first described Swiss acatalasemia [ 139 – 141 ]. The study on the fibroblast from Swiss acatalasemia patients suggests that structural mutations in the CAT gene are responsible for inactivation of catalase [ 142 ]. Goth, a Hungarian biochemist, first described Hungarian acatalasemia in 1992 after studying the disease in two Hungarian sisters. He found that the catalase activities in the blood of these two acatalasemic sisters were 4.4% and 6.7% of the reference catalase activity in the healthy population whereas the level of activity in hypocatalasemic patients was 38.9% [ 24 ]. Studies at his laboratory led Goth to suggest that mutations of the CAT gene and resultant structural changes in the catalase protein are responsible for Hungarian acatalasemia. This laboratory also reported that there was a risk of diabetes mellitus amongst the Hungarian acatalasemic family members though further biochemical and genetic analysis needs to be performed to validate the hypothesis that acatalasemic patients have more chance of developing diabetes mellitus [ 79 ]. There are generally four types of Hungarian acatalasemia which varies according to the (different) site of gene mutation in the DNA. The same is represented in Table 3 .

Four different types of Hungarian acatalasemia.

TypesPosition of mutationTypes of mutationResults of mutationEffect on catalaseReferences
Type AInsertion of GA at position 138 in exon 2 occurs which is responsible for the increase of the repeat number from 4 to 5Frame shift mutationCreates a TGA codon at position 134Lacks a histidine residue, an essential amino acid necessary for hydrogen peroxide binding[ ]
Type BInsertion of G at position 79 of exon 2Frame shift mutationGenerates a stop codon TGA at position 58A nonfunctional protein is produced[ ]
Type CA substitution mutation of G to A at position 5 in intron 7Splicing mutationNo change in peptide chainLevel of catalase protein expression is decreased[ , ]
Type DMutation of G to A at position 5 of exon 9Coding region mutationReplaces the arginine residue to histidine or cysteineLowering of catalase activity[ ]

4. Therapeutic Role of Catalase

Catalase is one of the most important antioxidant enzymes. As it decomposes hydrogen peroxide to innocuous products such as water and oxygen, catalase is used against numerous oxidative stress-related diseases as a therapeutic agent. The difficulty in application remains in delivering the catalase enzyme to the appropriate site in adequate amounts. Poly(lactic co-glycolic acid) nanoparticles have been used for delivering catalase to human neuronal cells, and the protection by these catalase-loaded nanoparticles against oxidative stress was evaluated [ 143 ]. It was observed that the efficiency of the encapsulation of catalase was very high with approximately 99% enzymatic activity of encapsulated catalase along with significantly sustained activity over a month. The nanoparticle-loaded catalase showed significant positive effect on neuronal cells preexposed to hydrogen peroxide reducing the hydrogen peroxide-mediated protein oxidation, DNA damage, mitochondrial membrane transition pore opening, and loss of membrane integrity. Thus, the study suggests that nanoparticle-loaded catalase may be used as a therapeutic agent in oxidative stress-related neurological diseases [ 143 ]. Similar research has been conducted using EUK 134 which is a class of synthetic superoxide dismutase/catalase mimetic as an effective therapeutic agent in stroke [ 144 ]. EUK 134 is a salen-manganese complex which has both high catalase and superoxide dismutase activity. It was concluded from these studies on the rat stroke model that EUK 134 may play a protective role in management of this disease.

Studies using Tat-CAT and 9Arg-CAT fusion proteins as therapeutic agents have also been carried out with encouraging results [ 145 ]. To study the effect of these fusion proteins under oxidative stress conditions, mammalian cell lines (HeLa, PC12) were transduced with purified fusion Tat-CAT and 9Arg-CAT protein and these cells were exposed to hydrogen peroxide. It was found that the viability of the transduced cells increased significantly. It was also observed that when the Tat-CAT and 9Arg-CAT fusion proteins were sprayed over animal skin, it could penetrate the epidermis and dermis layers of the skin. The fusion proteins transduced in mammalian cells were active enzymatically for over 60 h after which they became unstable. This study suggests that these fusion proteins can be potentially used as protein therapeutic agents in catalase-related disorders [ 145 ].

Amyotrophic lateral sclerosis (ALS) is one of the most common types of progressive and fatal neurological disorders which results in loss of motor neurons mostly in the spinal cord and also to some extent in the motor cortex and brain stem. Amongst the two distinct types of ALS, the familial form (FALS) accounts for 10% of all ALS cases and 15 to 20% of FALS cases are related to the SOD1 gene mutation, an antioxidant enzyme which scavenges the superoxide radical. In some of the FALS cases, it has been found that the mutation in the SOD1 gene is not linked to a lowered activity of SOD1. Rather, the mutated SOD1 has toxic properties with no lowering of the enzymatic activity. This mutated SOD1 protein reacts with some anomalous substrates such as hydrogen peroxide using it as a substrate and produces the most reactive hydroxyl radical which can severely damage important biomolecules [ 146 ]. Mutated SOD1 also has the potential to use peroxynitrite as an atypical substrate leading to the formation of 3-nitrotyrosine which results in the conversion of a functional protein into a nonfunctional one [ 147 ]. Catalase can reduce the hydrogen peroxide concentration by detoxifying it. Therapeutic approaches using putrescine-modified catalase in the treatment of FALS have also been attempted [ 148 ]. It was found that putrescine-catalase—a polyamine-modified catalase—delayed the progression of weakness in the FALS transgenic mouse model [ 148 ]. Thus, the delay in development of clinical weakness in FALS transgenic mice makes the putrescine-modified catalase a good candidate as a therapeutic agent in diseases linked with catalase anomaly. In this connection, it must be mentioned that the putrescine-modified catalase has been reported to exhibit an augmented blood-brain barrier permeability property while maintaining its activity comparable to that of native catalase with intact delivery to the central nervous system after parenteral administration [ 148 ]. Therefore, further studies with this molecule seem to be warranted.

Investigations using synthetic SOD-catalase mimetic, increase in the lifespan of SOD2 nullizygous mice along with recovery from spongiform encephalopathy, and alleviation of mitochondrial defects were observed [ 149 ]. These findings lead the authors to hypothesize that the SOD-catalase mimetic could be used as a potential therapy for different neurological diseases related to oxidative stress such as Alzheimer's disease and Parkinson's disease [ 149 ].

Studies using type 1 and type 2 diabetic mice models with 60-fold upregulated catalase expression showed amelioration in the functioning of the cardiomyocytes [ 150 ]. Cardiomyopathy is related to improper functioning of heart muscles where the muscles become enlarged, thick, or stiff. It can lead to irregular heartbeats or heart failure. Many diabetic patients suffer from cardiomyopathy with structural and functional anomalies of the myocardium without exhibiting concomitant coronary artery disease or hypertension [ 151 ].

As already discussed, catalase is interconnected to diabetes mellitus pathogenesis. It has been observed that a 60-fold increase of catalase activity could drastically reduce the usual features of diabetic cardiomyopathy in the mouse model [ 150 ]. Due to catalase overexpression, the morphological impairment of mitochondria and the myofibrils of heart tissue were prevented. The impaired cardiac contractility was also inhibited with decrease in the production of reactive oxygen species mediated by high glucose concentrations [ 150 ]. So this approach could be an effective therapeutic approach for the treatment of diabetic cardiomyopathy.

5. Future Perspective

This review summarizes a relation between catalase and the pathogenesis of some critical diseases such as diabetes, Parkinson's disease, acatalasemia, vitiligo, and Alzheimer's disease. An increase in focus on the role of catalase in the pathogenesis of oxidative stress-related diseases and its therapeutic approach is needed.

Catalase plays a significant role in hydrogen peroxide metabolism as a key regulator [ 28 , 29 , 152 – 154 ]. Some studies have also shown the involvement of catalase in controlling the concentration of hydrogen peroxide which is also involved in the signaling process [ 155 – 158 ]. Acatalasemia is a rare genetic disorder which is not as destructive as other diseases discussed here, but it could be a mediator in the development of other chronic diseases due to prolonged oxidative stress on the tissues.

We have also discussed the risk of type 2 diabetes mellitus among acatalasemic patients. But more research on the biochemical, molecular, and clinical aspects of the disease is necessary. There are many more questions about acatalasemia and its relation to other diseases which need to be answered. Therefore, further studies are needed to focus on catalase gene mutations and its relationship to acatalasemia and other diseases with decreased catalase activity so that the link can be understood more completely.

The therapeutic approaches using catalase needs more experimental validation so that clinical trials can be initiated. Use of catalase as a medicine or therapy may be a new and broad field of study. Any novel finding about therapeutic uses of catalase will have a huge contribution in medical science. Positive findings can direct towards its possible use for treatment of different oxidative stress-related diseases.

6. Conclusion

Catalase is one of the crucial antioxidant enzymes which plays an important role by breaking down hydrogen peroxide and maintaining the cellular redox homeostasis. Diabetes, Alzheimer's disease, Parkinson's disease, etc. are currently becoming common diseases. While there are many factors involved in the pathogenesis of these diseases, several studies from different laboratories have demonstrated that catalase has a relationship with the pathogenesis of these diseases. Research in this area is being carried out by many scientists at different laboratories exploring different aspects of these diseases, but with an ever-increasing aging population, much remains to be achieved. On the other hand, the potential of catalase as a therapeutic drug in the treatment of several oxidative stress-related diseases is not adequate and is still being explored. Additional research is needed to confirm if catalase may be used as a drug in the treatment of various age-related disorders.

Acknowledgments

The authors thank the Council of Scientific and Industrial Research (CSIR), India (Project no. 38(1343)/12/EMR-II) for EMR grants. AN is grateful to CSIR for the project fellowship and the University Grants Commission (UGC), India, for the non-NET fellowship.

Conflicts of Interest

The authors declare that they have no conflicts of interest.

Supplementary Materials

(Supplementary Figure 1). In module 1, ACOX1 (peroxisomal acyl coenzyme A oxidase), HSD17B4 (peroxisomal multifunctional enzyme), and HAO1 (hydroxyacid oxidase 1) are involved in the fatty acid oxidation pathway in the peroxisome while the protein DAO (D amino acid oxidase) is involved in the amino acid metabolism pathway in the peroxisome [ 4 – 6 ] (Supplementary Figure 1). All the components of module 1 are involved in different metabolic pathways. The proteins in module 2 are mainly involved in responses against oxidative stress. All the proteins have antioxidant activity except AKT1 (RAC-alpha serine-threonine protein kinase). AKT1 is a serine-threonine protein kinase which is involved in cell survival, metabolism, growth, and angiogenesis. All the proteins of both modules 1 and 2 including CAT have catalytic activity and are located in the lumen of intracellular organelles. SOD2 and AKT1 of module 2 including CAT were involved in the longevity regulating pathway and FOXO signaling pathway in mammals [ 4 – 6 ] (Supplementary Figures 2 and 3). But in multiple other species, SOD1 and SOD3 (superoxide dismutase 3) were also involved along with SOD2, AKT1, and CAT [ 4 – 6 ] (Supplementary Figure 4). Among the reactive species, hydrogen peroxide is freely diffusible and is relatively long-lived. It acts as a weak oxidizing as well as reducing agent; however, it is not very reactive, but it is the progenitor of many other reactive oxygen species (ROS). It has been demonstrated to oxidatively modify glyceraldehyde-3-phosphate dehydrogenase by oxidation of the labile essential thiol groups at the active site of this enzyme [ 2 ]. In most cellular injuries, this molecule is known to play an indirect role. One of the most important products is the formation of a more reactive free radical · OH radical in the presence of transition metal ions such as Fe 2+ by means of the Fenton reaction.

Information

  • Author Services

Initiatives

You are accessing a machine-readable page. In order to be human-readable, please install an RSS reader.

All articles published by MDPI are made immediately available worldwide under an open access license. No special permission is required to reuse all or part of the article published by MDPI, including figures and tables. For articles published under an open access Creative Common CC BY license, any part of the article may be reused without permission provided that the original article is clearly cited. For more information, please refer to https://www.mdpi.com/openaccess .

Feature papers represent the most advanced research with significant potential for high impact in the field. A Feature Paper should be a substantial original Article that involves several techniques or approaches, provides an outlook for future research directions and describes possible research applications.

Feature papers are submitted upon individual invitation or recommendation by the scientific editors and must receive positive feedback from the reviewers.

Editor’s Choice articles are based on recommendations by the scientific editors of MDPI journals from around the world. Editors select a small number of articles recently published in the journal that they believe will be particularly interesting to readers, or important in the respective research area. The aim is to provide a snapshot of some of the most exciting work published in the various research areas of the journal.

Original Submission Date Received: .

  • Active Journals
  • Find a Journal
  • Proceedings Series
  • For Authors
  • For Reviewers
  • For Editors
  • For Librarians
  • For Publishers
  • For Societies
  • For Conference Organizers
  • Open Access Policy
  • Institutional Open Access Program
  • Special Issues Guidelines
  • Editorial Process
  • Research and Publication Ethics
  • Article Processing Charges
  • Testimonials
  • Preprints.org
  • SciProfiles
  • Encyclopedia

diagnostics-logo

Article Menu

redox reaction and catalase activity experiment

  • Subscribe SciFeed
  • Recommended Articles
  • Google Scholar
  • on Google Scholar
  • Table of Contents

Find support for a specific problem in the support section of our website.

Please let us know what you think of our products and services.

Visit our dedicated information section to learn more about MDPI.

JSmol Viewer

Redox homeostasis and non-invasive assessment of significant liver fibrosis by shear wave elastography.

redox reaction and catalase activity experiment

1. Introduction

2. materials and methods, 4. discussions, 5. conclusions, author contributions, institutional review board statement, informed consent statement, data availability statement, conflicts of interest.

  • Khalifa, A.; Rockey, D.C. The utility of liver biopsy in 2020. Curr. Opin. Gastroenterol. 2020 , 36 , 184–191. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Venkatesh, S.K.; Torbenson, M.S. Liver fibrosis quantification. Abdom. Radiol. 2022 , 47 , 1032–1052. [ Google Scholar ] [ CrossRef ]
  • Tan, Z.; Sun, H.; Xue, T.; Gan, C.; Liu, H.; Xie, Y.; Yao, Y.; Ye, T. Liver Fibrosis: Therapeutic Targets and Advances in Drug Therapy. Front. Cell Dev. Biol. 2021 , 9 , 730176. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Elpek, G.O. Cellular and molecular mechanisms in the pathogenesis of liver fibrosis: An update. World J. Gastroenterol. 2014 , 20 , 7260–7276. [ Google Scholar ] [ CrossRef ]
  • Gressner, A.M.; Weiskirchen, R.; Breitkopf, K.; Dooley, S. Roles of TGF-beta in hepatic fibrosis. Front. Biosci. 2002 , 7 , d793–d807. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Gressner, O.A.; Weiskirchen, R.; Gressner, A.M. Biomarkers of liver fibrosis: Clinical translation of molecular pathogenesis or based on liver-dependent malfunction tests. Clin. Chim. Acta 2007 , 381 , 107–113. [ Google Scholar ] [ CrossRef ]
  • Guechot, J.; Laudat, A.; Loria, A.; Serfaty, L.; Poupon, R.; Giboudeau, J. Diagnostic accuracy of hyaluronan and type III procollagen amino-terminal peptide serum assays as markers of liver fibrosis in chronic viral hepatitis C evaluated by ROC curve analysis. Clin. Chem. 1996 , 42 , 558–563. [ Google Scholar ] [ CrossRef ]
  • Liu, T.; Wang, X.; Karsdal, M.A.; Leeming, D.J.; Genovese, F. Molecular serum markers of liver fibrosis. Biomark. Insights 2012 , 7 , 105–117. [ Google Scholar ] [ CrossRef ]
  • Berzigotti, A.; Ashkenazi, E.; Reverter, E.; Abraldes, J.G.; Bosch, J. Non-invasive diagnostic and prognostic evaluation of liver cirrhosis and portal hypertension. Dis. Mark. 2011 , 31 , 129–138. [ Google Scholar ] [ CrossRef ]
  • Papastergiou, V.; Tsochatzis, E.; Burroughs, A.K. Non-invasive assessment of liver fibrosis. Ann. Gastroenterol. 2012 , 25 , 218–231. [ Google Scholar ]
  • Lurie, Y.; Webb, M.; Cytter-Kuint, R.; Shteingart, S.; Lederkremer, G.Z. Non-invasive diagnosis of liver fibrosis and cirrhosis. World J. Gastroenterol. 2015 , 21 , 11567–11583. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Okuda, M.; Li, K.; Beard, M.R.; Showalter, L.A.; Scholle, F.; Lemon, S.M.; Weinman, S.A. Mitochondrial injury, oxidative stress, and antioxidant gene expression are induced by hepatitis C virus core protein. Gastroenterology 2002 , 122 , 366–375. [ Google Scholar ] [ CrossRef ]
  • Abrahamovych, M.; Abrahamovych, O.; Fayura, O.; Tolopko, S. Relation between redox homeostasis blood parameters in cirrhotic patients and endothelial dysfunction development. Min. Gastroenterol. Dietol. 2020 , 66 , 98–105. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Gerber, L.; Kasper, D.; Fitting, D.; Knop, V.; Vermehren, A.; Sprinzl, K.; Hansmann, M.L.; Herrmann, E.; Bojunga, J.; Albert, J.; et al. Assessment of liver fibrosis with 2-D shear wave elastography in comparison to transient elastography and acoustic radiation force impulse imaging in patients with chronic liver disease. Ultras Med. Biol. 2015 , 41 , 2350–2359. [ Google Scholar ] [ CrossRef ]
  • Virarkar, M.; Morani, A.C.; Taggart, M.W.; Bhosale, P. Liver Fibrosis Assessment. Semin. Ultrasound CT MRI 2021 , 42 , 381–389. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Branchi, F.; Conti, C.B.; Baccarin, A.; Lampertico, P.; Conte, D.; Fraquelli, M. Non-invasive assessment of liver fibrosis in chronic hepatitis B. World J. Gastroenterol. 2014 , 20 , 14568–14580. [ Google Scholar ] [ CrossRef ]
  • Boursier, J.; de Ledinghen, V.; Leroy, V.; Anty, R.; Francque, S.; Salmon, D.; Lannes, A.; Bertrais, S.; Oberti, F.; Fouchard-Hubert, I.; et al. A stepwise algorithm using an at-a-glance first-line test for the non-invasive diagnosis of advanced liver fibrosis and cirrhosis. J. Hepatol. 2017 , 66 , 1158–1165. [ Google Scholar ] [ CrossRef ]
  • Bota, S.; Sporea, I.; Sirli, R.; Focsa, M.; Popescu, A.; Danila, M.; Strain, M. Can ARFI elastography predict the presence of significant esophageal varices in newly diagnosed cirrhotic patients? Ann. Hepatol. 2012 , 11 , 519–525. [ Google Scholar ] [ CrossRef ]
  • Mózes, F.E.; Lee, J.A.; Selvaraj, E.A.; Jayaswal, A.N.A.; Trauner, M. Diagnostic accuracy of non-invasive tests for advanced fibrosis in patients with NAFLD: An individual patient data meta-analysis. Gut 2021 , 71 , 1006–1019. [ Google Scholar ] [ CrossRef ]
  • Sebastiani, G.; Halfon, P.; Castera, L.; Mangia, A.; Di Marco, V.; Pirisi, M.; Voiculescu, M.; Bourliere, M.; Alberti, A. Comparison of three algorithms of non-invasive markers of fibrosis in chronic hepatitis C. Aliment. Pharmacol. Ther. 2012 , 35 , 92–104. [ Google Scholar ] [ CrossRef ]
  • Piscaglia, F.; Salvatore, V.; Mulazzani, L.; Cantisani, V.; Schiavone, C. Ultrasound Shear Wave Elastography for Liver Disease. A Critical Appraisal of the Many Actors on the Stage. Ultraschall Med. 2016 , 37 , 1–5. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Ferraioli, G.; Parekh, P.; Levitov, A.B.; Filice, C. Shear Wave Elastography for Evaluation of Liver Fibrosis. J. Ultrasound Med. 2014 , 33 , 197–203. [ Google Scholar ] [ CrossRef ]
  • Kleiner, D.; Szilvas, A.; Szentmihalyi, K.; Sule, K.; Blazovics, A. Changes of erythrocyte element status of colectomysed cancerous patients: Retrospective study. J. Trace Elem. Med. Biol. 2016 , 33 , 8–13. [ Google Scholar ] [ CrossRef ]
  • Ellman, G.L.; Lysko, H. Disulfide and sulfhydryl compounds in TCA extracts of human blood and plasma. J. Lab. Clin. Med. 1967 , 70 , 518–527. [ Google Scholar ]
  • Hatano, T.; Kagawa, H.; Yasuhara, T.; Okuda, T. Two new flavonoids and other constituents in licorice root: Their relative astringency and radical scavenging effects. Chem. Pharm. Bull. 1988 , 36 , 2090–2097. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Blázovics, A.; Sárdi, É. Methodological repertoire development to study the effect of dietary supplementation in cancer therapy. Microchem. J. 2018 , 136 , 121–127. [ Google Scholar ] [ CrossRef ]
  • Hauck, A.K.; Huang, Y.; Hertzel, A.V.; Bernlohr, D.A. Adipose oxidative stress and protein carbonylation. J. Biol. Chem. 2019 , 294 , 1083–1088. [ Google Scholar ] [ CrossRef ]
  • Luo, Q.T.; Zhu, Q. Diagnostic Performance of Transient Elastography Versus Two-Dimensional Shear Wave Elastography for Liver Fibrosis in Chronic Viral Hepatitis: Direct Comparison and a Meta-Analysis. BioMed Res. Int. 2022 , 2022 , 1960244. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Parola, M.; Pinzani, M. Liver fibrosis: Pathophysiology, pathogenetic targets and clinical issues. Mol. Asp. Med. 2019 , 65 , 37–55. [ Google Scholar ] [ CrossRef ]
  • Sigrist, R.M.S.; Liau, J.; Kaffas, A.E.; Chammas, M.C.; Willmann, J.K. Ultrasound Elastography: Review of Techniques and Clinical Applications. Theranostics 2017 , 7 , 1303–1329. [ Google Scholar ] [ CrossRef ]
  • Stasi, C.; Milani, S. Evolving strategies for liver fibrosis staging: Non-invasive assessment. World J. Gastroenterol. 2017 , 23 , 191–196. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Cichoz-Lach, H.; Michalak, A. Oxidative stress as a crucial factor in liver diseases. World J. Gastroenterol. 2014 , 20 , 8082–8091. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Crosas-Molist, E.; Fabregat, I. Role of NADPH oxidases in the redox biology of liver fibrosis. Redox Biol. 2015 , 6 , 106–111. [ Google Scholar ] [ CrossRef ]
  • Ozturk, A.; Olson, M.C.; Samir, A.E.; Venkatesh, S.K. Liver fibrosis assessment: MR and US elastography. Abdom. Radiol. 2022 , 47 , 3037–3050. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Fu, J.; Wu, B.; Wu, H.; Lin, F.; Deng, W. Accuracy of real-time shear wave elastography in staging hepatic fibrosis: A meta-analysis. BMC Med. Imaging 2020 , 20 , 16. [ Google Scholar ] [ CrossRef ]
  • Barr, R.G.; Wilson, S.R. Update to the Society of Radiologists in Ultrasound Liver Elastography Consensus Statement. Radiology 2020 , 296 , 263–274. [ Google Scholar ] [ CrossRef ]
  • Zayadeen, A.R.; Hijazeen, S.; Smadi, M.; Fayyad, L.; Halasa, M.; AlQusous, S.; AlRabadi, O.; Hijazeen, R.; Ajlouni, Y.; Tulenko, K.; et al. Comparing shear wave elastography with liver biopsy in the assessment of liver fibrosis at King Hussein Medical Center. Egypt. Liver J. 2022 , 12 , 24. [ Google Scholar ] [ CrossRef ]
  • Ferraioli, G.; Wong, V.W.-S.; Castera, L.; Berzigotti, A.; Sporea, I.; Dietrich, C.F.; Choi, B.I.; Wilson, S.R.; Kudo, M.; Barr, R.G. Liver Ultrasound Elastography: An Update to the World Federation for Ultrasound in Medicine and Biology Guidelines and Recommendations. Ultrasound Med. Biol. 2018 , 44 , 2419–2440. [ Google Scholar ] [ CrossRef ]
  • Stern, C.; Ngo, A.; Villela-Nogueira, C.; Thabut, D.; Ratziu, V. A Simple and Reliable 2D-Shear Wave Elastography and UltraSound Coefficient Attenuation Parameter Technique in Chronic Liver Diseases. Dig. Dis. Sci. 2024 , 69 , 2648–2654. [ Google Scholar ] [ CrossRef ]
  • Haghgoo, S.M.; Sharafi, H.; Alavian, S.M. Serum cytokines, adipokines and ferritin for non-invasive assessment of liver fibrosis in chronic liver disease: A systematic review. Clin. Chem. Lab. Med. 2018 , 57 , 577–610. [ Google Scholar ] [ CrossRef ]
  • Abu-Tair, L.; Doron, S.; Mahamid, M.; Amer, J.; Safadi, R. Leptin modulates lymphocytes’ adherence to hepatic stellate cells is associated with oxidative status alterations. Mitochondrion 2013 , 13 , 473–480. [ Google Scholar ] [ CrossRef ]
  • Makri, E.S.; Evripidou, K.; Polyzos, S.A. Circulating leptin in patients with nonalcoholic fatty liver disease-related liver fibrosis: A systematic review and a meta-analysis. J. Gastroenterol. Hepatol. 2024 , 39 , 806–817. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Udomsinprasert, W.; Honsawek, S.; Poovorawan, Y. Adiponectin as a novel biomarker for liver fibrosis. World J. Hepatol. 2018 , 10 , 708–718. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Angin, Y.; Arslan, N.; Kuralay, F. Leptin-to-adiponectin ratio in obese adolescents with nonalcoholic fatty liver disease. Turk. J. Pediatr. 2014 , 56 , 259–266. [ Google Scholar ] [ PubMed ]
  • Kamada, Y.; Tamura, S.; Kiso, S.; Matsumoto, H.; Saji, Y.; Yoshida, Y.; Fukui, K.; Maeda, N.; Nishizawa, H.; Nagaretani, H.; et al. Enhanced carbon tetrachloride-induced liver fibrosis in mice lacking adiponectin. Gastroenterology 2003 , 125 , 1796–1807. [ Google Scholar ] [ CrossRef ]
  • Saxena, N.K.; Anania, F.A. Adipocytokines and hepatic fibrosis. Trends Endocrinol. Metab. 2015 , 26 , 153–161. [ Google Scholar ] [ CrossRef ]
  • Dong, Z.; Su, L.; Esmaili, S.; Iseli, T.J.; Ramezani-Moghadam, M.; Hu, L.; Xu, A.; George, J.; Wang, J. Adiponectin attenuates liver fibrosis by inducing nitric oxide production of hepatic stellate cells. J. Mol. Med. 2015 , 93 , 1327–1339. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Tontikidou, C.; Makri, E.S.; Evripidou, K.; Goulis, D.G.; Goulas, A.; Polyzos, S.A. Circulating adiponectin in patients with nonalcoholic fatty liver disease-related liver fibrosis: A systematic review and a meta-analysis. J. Gastroenterol. Hepatol. 2022 , 37 , 1853–1864. [ Google Scholar ] [ CrossRef ]
  • Fitzpatrick, E.; Dhawan, A. Noninvasive biomarkers in non-alcoholic fatty liver disease: Current status and a glimpse of the future. World J. Gastroenterol. 2014 , 20 , 10851–10863. [ Google Scholar ] [ CrossRef ]
  • Neuman, M.G.; Maor, Y.; Nanau, R.M.; Melzer, E.; Mell, H.; Opris, M.; Cohen, L.; Malnick, S. Alcoholic Liver Disease: Role of Cytokines. Biomolecules 2015 , 5 , 2023–2034. [ Google Scholar ] [ CrossRef ]
  • Jung, Y.K.; Yim, H.J. Reversal of liver cirrhosis: Current evidence and expectations. Korean J. Intern. Med. 2017 , 32 , 213–228. [ Google Scholar ] [ CrossRef ]
  • Arauz, J.; Ramos-Tovar, E.; Muriel, P. Redox state and methods to evaluate oxidative stress in liver damage: From bench to bedside. Ann. Hepatol. 2016 , 15 , 160–173. [ Google Scholar ] [ CrossRef ]
  • Milito, A.; Brancaccio, M.; D’Argenio, G.; Castellano, I. Natural Sulfur-Containing Compounds: An Alternative Therapeutic Strategy against Liver Fibrosis. Cells 2019 , 8 , 1356. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Richter, K.; Konzack, A.; Pihlajaniemi, T.; Heljasvaara, R.; Kietzmann, T. Redox-fibrosis: Impact of TGFβ1 on ROS generators, mediators and functional consequences. Redox Biol. 2015 , 6 , 344–352. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Karanjia, R.N.; Crossey, M.M.; Cox, I.J.; Fye, H.K.; Njie, R.; Goldin, R.D.; Taylor-Robinson, S.D. Hepatic steatosis and fibrosis: Non-invasive assessment. World J. Gastroenterol. 2016 , 22 , 9880–9897. [ Google Scholar ] [ CrossRef ]

Click here to enlarge figure

DimensionF0–F1F2F3F4F2–4
Leptin (ng/mL)4.88
±
7.23
5.25
±
5.60
4.51
±
5.59
6.20
±
7.23
5.45
±
5.94
Adiponectin (pg/mL)12,374.68
±
13,658.2
9560.93
±
6356.05
24,665
±
21,830.25
16,810
±
12,070.07
15,629.7
±
13,801.38
Interleukin-6 (pg/mL)1.36
±
1.19
1.69
±
1.65
2.21
±
2.34
2.45
±
2.93
2.12
±
2.28
TNF-α (pg/mL)78.91
±
72.89
155.47
±
151.01
41.07
±
28.06
63.65
±
57.59
94.34
±
109.33
The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

Egresi, A.; Blázovics, A.; Lengyel, G.; Tóth, A.G.; Csongrády, B.; Jakab, Z.; Hagymási, K. Redox Homeostasis and Non-Invasive Assessment of Significant Liver Fibrosis by Shear Wave Elastography. Diagnostics 2024 , 14 , 1945. https://doi.org/10.3390/diagnostics14171945

Egresi A, Blázovics A, Lengyel G, Tóth AG, Csongrády B, Jakab Z, Hagymási K. Redox Homeostasis and Non-Invasive Assessment of Significant Liver Fibrosis by Shear Wave Elastography. Diagnostics . 2024; 14(17):1945. https://doi.org/10.3390/diagnostics14171945

Egresi, Anna, Anna Blázovics, Gabriella Lengyel, Adrienn Gréta Tóth, Barbara Csongrády, Zsuzsanna Jakab, and Krisztina Hagymási. 2024. "Redox Homeostasis and Non-Invasive Assessment of Significant Liver Fibrosis by Shear Wave Elastography" Diagnostics 14, no. 17: 1945. https://doi.org/10.3390/diagnostics14171945

Article Metrics

Further information, mdpi initiatives, follow mdpi.

MDPI

Subscribe to receive issue release notifications and newsletters from MDPI journals

  • Open access
  • Published: 29 August 2024

Neural acetylcholinesterase and monoamine oxidase deregulation during streptozotocin-induced behavioral, metabolic and redox modification in Nauphoeta cinerea

  • Opeyemi B. Ogunsuyi 1 , 2 , 3 ,
  • Olawande C. Olagoke 4 , 5 , 6 ,
  • Mayokun E. Famutimi 2 , 3 ,
  • Damilola M. Olatunde 3 , 7 ,
  • Diogo O. G. Souza 8 ,
  • Ganiyu Oboh 3 , 7 ,
  • Nilda V. Barbosa 1 &
  • João B.T. Rocha 1 , 8  

BMC Neuroscience volume  25 , Article number:  42 ( 2024 ) Cite this article

Metrics details

Genetic and environmental factors have been linked with neurodegeneration, especially in the elderly. Yet, efforts to impede neurodegenerative processes have at best addressed symptoms instead of underlying pathologies. The gap in the understanding of neuro-behavioral plasticity is consistent from insects to mammals, and cockroaches have been proven to be effective models for studying the toxicity mechanisms of various chemicals. We therefore used head injection of 74 and 740 nmol STZ in Nauphoeta cinerea to elucidate the mechanisms of chemical-induced neurotoxicity, as STZ is known to cross the blood-brain barrier. Neurolocomotor assessment was carried out in a new environment, while head homogenate was used to estimate metabolic, neurotransmitter and redox activities, followed by RT-qPCR validation of relevant cellular signaling. STZ treatment reduced the distance and maximum speed travelled by cockroaches, and increased glucose levels while reducing triglyceride levels in neural tissues. The activity of neurotransmitter regulators – AChE and MAO was exacerbated, with concurrent upregulation of glucose sensing and signaling, and increased mRNA levels of redox regulators and inflammation-related genes. Consequently, STZ neurotoxicity is conserved in insects, with possible implications for using N. cinerea to target the multi-faceted mechanisms of neurodegeneration and test potential anti-neurodegenerative agents.

Peer Review reports

Introduction

Advances in medicine and technology have increased the average life expectancy, especially in the global north, but ageing is strongly linked with neurological disorders that affect the brain and some other regions of the nervous system, including neurodegenerative diseases that are often irreversible, progressive, and characterized by diminished cognitive abilities, muscle weakness and locomotor difficulties [ 1 ]. Regardless of the cause, neurodegenerative diseases often portray features of neuronal loss or damage, and the type of disorder depends on the extent and site of damage [ 2 ]. However, the heterogeneous nature of population-based genetics and the complexities in inheritance patterns increase the difficulty of pathway analysis of human data, making it important to exploit models that may explain the pathophysiological mechanisms driving changes in protein structure and function across multiple generations [ 3 , 4 ].

Insect models like Drosophila melanogaster and Nauphoeta cinerea have been insightful in depicting fundamental propositions like the chromosomal theory of inheritance or mechanistic patterns like xenobiotic-related neurotoxicity [ 5 , 6 , 7 ]. The lobster cockroach has a high fecundity rate, with each female producing about 20 oothecas that house 26 to 40 eggs [ 8 ], it is cheap and easy to maintain in a laboratory setting, and the neural tissues are more accessible given the comparative larger size to the Drosophila. Importantly, although the insect open circulatory system is different from the mammalian closed circulatory system, the brain in insects and mammals is protected from fluctuations in solute concentration by the hemolymph-brain barrier and blood-brain barrier respectively [ 9 ], and metabolic pathways are conserved from insects to mammals [ 10 , 11 , 12 ]. For example, ablation of insect insulin-producing cells (IPC) creates similar phenotypes to those seen in mammals with beta cell damage, even though insect IPCs are found in the brain and mammalian beta cells are found in the pancreas [ 13 ].

The tropism and toxicity of streptozotocin (STZ) towards the mammalian beta cells induces DNA alkylation [ 14 ] and nicotinamide adenine dinucleotide depletion [ 15 ] with a consequent disruption in glucose energy metabolism. Likewise, intracerebroventricular (ICV) administration of STZ in rodents induces insulin resistance in the brain that results in hyperphosphorylation of tau proteins and aggregation of Aβ in meningeal vessels, which causes neural and behavioral changes that are reminiscent of sporadic Alzheimer disease [ 16 ]. Moreover, STZ is mutagenic to the mosquito cell [ 17 ], and the intraabdominal administration of STZ in the lobster cockroach alters brain glucose metabolism, including upregulation of glucose transporter expression, increased glucose absorption into neural tissues and deregulation of redox balance [ 12 , 18 , 19 ]. Similar phenotypes are therefore seen across insects and rodents that are exposed to STZ.

Nonetheless, several knowledge gaps remain in the understanding of the neurodegenerative process, which may explain why current management plans are heavily focused on alleviating symptoms, instead of providing cures. For example, the limited understanding of neurobehavioral plasticity persists from insects to mammals. Given the homology between insect and mammalian signaling [ 12 , 20 ], and the reports of the effectiveness of using cockroaches to illustrate the interaction of chemicals with living systems [ 21 , 22 , 23 , 24 , 25 ], we used Nauphoeta cinerea to explore chemical-induced neurotoxicity, to provide insights into the mechanisms driving neurodegeneration, while advancing the prospect of replacing, reducing and refining (3Rs) animal use in biomedical research.

Materials and methods

Streptozotocin (Sc-200719 A) was sourced from Santa Cruz Biotechnologies, Germany. Chemical reagents such as acetylthiocholine iodide, reduced glutathione, and ferrous sulphate, were procured from Sigma Aldrich Co. (St Louis, Missouri, USA). Trichloroacetic acid (TCA) and sodium acetate were sourced from Sigma Aldrich (Steinheim, Germany). Acetic acid, hydrochloric acid, aluminum chloride, and potassium acetate, were obtained from BDH Chemicals Ltd., (Poole, England). Except stated otherwise, all other chemicals and reagents were of analytical grades and the water was glass distilled.

N. Cinerea stock culture

N. cinerea was raised at the Biochemistry Department from UFSM, RS, Brazil. The cockroaches were maintained and reared on a commercial dog chow at constant temperature at constant temperature (24 ± 3 °C) and humidity (57–75%).

Experimental design

40 size-matched cockroaches (including male and female) were randomly selected per study group and were housed in plastic containers. The inner edges of the housing containers were lubricated to prevent the cockroaches from crawling out, after which the cockroaches were weighed (mg) using a weighing scale. There were three groups in total: the control group, the 74 nm STZ injected group, and the 740 nm STZ injected group. A preliminary study was done to test a sham injection of 0.8% NaCl, but there was no difference in the biochemical properties of untreated and sham-injected cockroaches, hence untreated cockroaches were used as the control group. Streptozotocin concentrations were based on previously reported pilot studies in Nauphoeta cinerea that were extrapolated from data on the injection of β- cell cytotoxic agents in moths and rodents [ 18 , 26 , 27 ].

All groups were anaesthetized using ice. The period of anaesthesia was about 5 min to reduce the agility of the cockroaches during induction. Care was taken not to keep the cockroaches under anaesthesia for longer than 5 min to prevent death before induction. Single-use insulin syringes were then used to administer 20 µL STZ (74 nm and 740 nm respectively) directly to the central axes of the cockroach heads, and the gross toxicity of STZ was determined in terms of insect survival [ 18 ]. 1.7 g of flour, 0.2 g of sugar, 0.05 g of casein, 0.01 g of salt and 0.04 g of milk were weighed using a High Precision Balance and were thoroughly but carefully mixed using a Vortex mixer [ 18 ]. The cockroaches had access to water and feed ad libitum and were observed for 7 days, after which behavioural profiles were assessed.

Neurolocomotor assessment

The cockroaches were carefully transferred into a white plastic box (a new environment) 19 cm in length, 12.5 cm in width and 5 cm in height and their behavior was filmed during a 10 min trial period using an overhead-mounted webcam. Behavioural endpoints of locomotor activity such as total distance travelled, maximum speed, total time immobile, and total time in periphery were analyzed from the video files using the video-tracking software; ANY-maze 6.0, Steolting, CO, USA, as earlier described [ 28 , 29 ].

Biochemical analysis

The cockroach heads were excised using surgical blades. Three heads per vial ( n  = 5) were then weighed, homogenized in 0.1 M phosphate buffer, pH 7.4, and centrifuged at 2500 g for 10 min at 4 o C. The resulting tissue was used to carry out biochemical assays. Total protein was estimated by applying 2 µL tissue homogenate to the sensor of a NanoDrop 2000 spectrophotometer (ThermoFisher Scientific Inc, USA) and reading the absorbance at 280 nm as earlier described [ 30 , 31 ].

Head glucose and triglyceride content

Glucose content was determined using a glucose oxidase kit (Lot: 07135A0217) according to the manufacturer’s protocol (Quimiglic-Ox, Brazil). In brief, 196 µL glucose reagent and 4 µL tissue was added to the sample wells, while the blank well was made up of 196 µL glucose reagent and 4 µL distilled water. The standard well was made up of 4 µL glucose standard and 196 µL glucose reagent. The mixture was then incubated for 10 min at 37 o C, after which it was read at 505 nm in a BIO-RAD Microplate reader.

Triglyceride content was determined using a triglyceride kit according to the manufacturer’s protocol (Biotecnica, Brazil) In brief, 196 µL triglyceride reagent and 4 µL tissue was added to the sample wells, while the blank well was made up of 196 µL triglyceride reagent and 4 µL distilled water. The standard well was made up of 4 µL triglyceride standard and 196 µL triglyceride reagent. The mixture was then incubated for 10 min at 37 o C, after which it was read at 505 nm in a BIO-RAD Microplate reader [ 18 ].

Acetylcholinesterase and monoamine oxidase activity

Acetylcholinesterase activity was carried out in an assay medium consisting of (at final concentration) 10 mM phosphate buffer (pH 7.4), 1 mM 5,5-dithio-bis (2-nitrobenzoic) acid (DTNB), and 0.8 mM acetylthiocholine iodide, 30 µL of tissue sample and the total reaction volume made up to 200 µL with distilled water. The mixture was then read using a SpectraMax Microplate Spectrophotometer at 412 nm (15s intervals for 30 min). The AChE activity was thereafter calculated and expressed as mmolAcSch/h/mg protein [ 21 ].

Monoamine oxidase (MAO) activity was determined as previously reported [ 32 , 33 ]. The reaction consisted (final concentration) of 72 mM potassium phosphate buffer (pH 7.4), 0. 5 mM benzylamine, 50 µl of tissue homogenate, and 50 µl of distilled water. This was followed by incubating the mixture for 30 min at 25 degrees Celsius and adding 300 µl of 10% perchloric acid (5.2% of total reaction volume). The mixture was thereafter centrifuged at 1,500 g for 10 min. The MAO activity was monitored at 280 nm and expressed as mmol/mg protein.

Oxidative stress assay

Total reactive oxygen and nitrogen species were assayed in a reaction which included 5 µM 2′, 7′-dichlorofluorescein diacetate (2, 7-DCFDA), 5 µL of homogenate, and 75 mM potassium phosphate buffer (pH 7.4). After that, for 30 min at 30-second intervals, the fluorescence emission was observed using a Spectra Max spectrofluorometer (excitation = 480 nm; emission = 525 nm). The results were represented as a change in fluorescence intensity per minute. [ 21 , 23 , 34 ].

The reactive oxygen species (ROS) level in the tissue homogenates was also estimated as H 2 O 2 equivalent according to the method of Hayashi et al. [ 35 ] as previously reported by Oboh et al. [ 36 ]. In brief, 5 µl of tissue homogenate and 57 mM sodium acetate buffer (pH 4.8) were incubated for 5 min at 37 degrees Celsius. Thereafter, 2.5 mg/mL of n-n-diethyl-para-phenylenediamine (DEPPD) and 1.8 µM of ferrous sulphate solution were added to the mixture. The absorbance was measured at 505 nm using a spectrophotometer. ROS levels was estimated from an H 2 O 2 standard calibration curve and expressed as Unit/mg protein, where 1 unit = 1 mg H 2 0 2 /L.

The lipid peroxidation was estimated using a mixture of 50 µL tissue, 150 µL 8.1% SDS, 250 µL 20% acetic acid in hydrogen chloride (p.H 3.4) and 250 µL 0.6% TBA for the sample test tubes; and 50 µL distilled water, 150 µL 8.1% SDS, 250 µL 20% acetic acid in hydrogen chloride (p.H 3.4) and 250 µL 0.6% TBA in the blank test tube. The mixture was then incubated for 1 h at 95 o C, after which it was read at 532 nm in a visible Spectrophotometer [ 37 , 38 ].

Antioxidant and detoxification activity

Using Ellman et al.‘s method [ 39 ], the amount of total thiol content in the tissue homogenate was measured. The reaction mixture included 20 µL of tissue homogenate, 0.5 mM 5,5’-dithiobis-(2-nitrobenzoic acid) (DTNB), and 85 mM potassium phosphate buffer (pH 7.4). The total reaction volume was 200 µL. The same quantity of tissue was added to reaction blanks, but DTNB was not present. After that, the sample was allowed to incubate for 30 min at room temperature, and the absorbance was measured at 412 nm. After that, the total thiol content was determined and represented as (mmol/mg protein).

Glutathione-S-transferase activity was assayed according to the method of Habig and Jakoby with slight modifications, using 1-chloro-2,4-dinitrobenzene (CDNB) as a substrate [ 40 , 41 ]. This assay consists of 100 µL of tissue homogenate, 1. mM ethylenediaminetetraacetic acid, 0.80 mM chloro-2, 4-dinitrobenzene (CDNB), 3.20 mM glutathione as substrate and 70 mM potassium phosphate buffer (pH 7.0). After ten minutes, the GST activity was measured using a spectrophotometer (Spectra Max) plate reader at 340 nm. The result was expressed as unit/mg protein where 1 unit = µmol/ml/min/mg protein.

Real-time polymerase chain reaction

Our Nauphoeta cinerea transcriptome [ 12 , 42 ] was queried with the existing insect gene sequences and the ensuing Nauphoeta gene sequences were used to design primers as earlier described [ 18 , 19 ] and shown in Table  1 . A reverse transcription-quantitative real-time polymerase chain reaction was used to evaluate the mRNA expressions. Fastzol™ reagent (Lot: 211711 by Quatro G Biotecnologia, Brazil) was used to extract the total RNA from the cockroach head. The concentration of RNA in the samples was measured with a Nanodrop400A™, thereafter, it was visualized in a 1.5% agarose gel ad then treated with DNase I (M0303S by New England Biolabs). The High-capacity cDNA Reverse Transcription Kit (Lot: 4368814 by Applied Biosystems) was used to synthesize 1 µg of cDNA in a thermal cycler (BIO-RAD, USA). All levels of expression were normalized to TBP. Real-time PCR was carried out using a QuantStudio 3 RT-qPCR System (ThermoFisher Scientific, USA). Each well contained a final volume of 20 µL, 2.5 ng/µL cDNA, 10 µL MasterMix PCR Tempo Real - SYBR Green/ROX (Lot: P242401001 by Quatro G Biotecnologia, Brazil) and 0.25 µM of forward and reverse primers (Table  1 ). The PCR entailed 40 thermal cycles of 15 s at 94 °C, 10 s at 60 °C, and 30 s at 72 °C, and a melt curve stage with 1 cycle of 94 °C for 10 s, 55 °C for 1 min and 94 °C for 15 s. Primer efficiency was determined using a five-point dilution of a pool of samples [ 19 ], and the SYBR fluorescence was measured using the StepOne™ design and analysis program [ 43 ]. The experiments were performed in triplicates, and the 2 −ΔΔCT approach [ 44 ] was applied to calculate gene expression levels.

Statistical analysis

The results were expressed as mean ± standard deviation (SD) and were analyzed by one-way Analysis of Variance (ANOVA) after the Shapiro-Wilk Test for normality. Post hoc analysis was carried out with Tukey’s multiple comparisons test. All statistical analysis was carried out using the software Graph pad PRISM (V.8.0), and significance was set at p ≤ 0.05.

Survival and behavioural monitoring

We recorded the trends in mortality rate, as well as behavioural responses to 74 and 740 nmol head STZ injection and found significantly diminished survival in the treated group compared with the control group. Expectedly, there was higher mortality in the 740 nmol STZ treated group compared with the 74 nmol STZ treated group (Fig.  1 ). Head STZ injection also disrupted N. Cinerea motor and exploratory activities (Fig.  1 A & B ), as there was a significant reduction in the total distance travelled [F (2,21), = 12.66; P  = 0.0002; Fig.  2 C] and maximum speed [F (2,21), = 29.61; P  < 0.0001; Fig.  2 D] of the treated groups in relation to control. Conversely, the treated group spent significantly more time immobile [F (2,16), = 18.3; P  < 0.0001; Fig.  2 E] and in the periphery [F (2,17), = 6.9; P  < 0.005; Fig.  2 F] compared with the control group. However, the behavioral endpoints were not significantly different across the treatment (74 and 740 nmol STZ) groups.

figure 1

7-day Kaplan-Meier survival analysis of cockroaches after single dose head STZ injection ( N  = 40). Log-rank (Mantel-Cox) test showed significant ( P  = 0.0003) reduction in survival (increased mortality) of 74 and 740 nM STZ-treated cockroaches compared with control

figure 2

Motor and exploratory profile of cockroaches, 7 days after head STZ injection. n  = 8. 8-minute video recordings of ( A ) track plot and ( B ) heat map were analysed with the ANY-maze (Stoelting CO, USA) software. One-way ANOVA with Tukey’s multiple comparisons test indicated a significant decrease in ( C ) total distance travelled and ( D ) maximum speed, as well as a significant increase in ( E ) total time immobile and ( F ) total time in periphery in cockroaches exposed to 74 and 740 nmol STZ head injection, compared with cockroaches in the control group. All values are mean ± SD. * indicates a significant difference from control

Biochemical analyses

Head glucose levels were significantly increased [F (2,9), = 83.77; P  < 0.0001; Fig.  3 A], while triglyceride levels were significantly reduced [F (2,9), = 483.0; P  < 0.0001; Fig.  3 B], and acetylcholinesterase (AChE) [F (2,9), = 91.89; P  < 0.0001; Fig.  4 A] was significantly increased in the 74 and 740 nmol STZ-treated groups compared with control groups. 74 nmol STZ treatment was also significantly different from 740 nmol STZ treatment in the glucose, triglyceride and acetylcholinesterase assays. Monoamine oxidase (MAO) was significantly increased [F (2,9), = 91.89; P  < 0.0001; Fig.  4 B] only in the 74 nmol STZ treated cockroaches compared with control.

figure 3

Sugar and lipid levels, 7 days after head STZ injection in cockroaches. n  = 5. One-way ANOVA with Tukey’s multiple comparisons test indicated a significant increase in ( A ) Glucose levels, and a significant decrease in ( B ) Triglyceride levels in neural tissues of cockroaches exposed to 74 and 740 nM STZ head injection, compared with cockroaches in the control group. All values are mean ± SD. * indicates a significant difference from control; # indicates significant differences from 74 nM STZ injection

figure 4

Increased neurotransmitter regulator activity, 7 days after head STZ injection in cockroaches. n  = 5. One-way ANOVA with Tukey’s multiple comparisons test indicated a significant increase in ( A ) AChE activity and ( B ) MAO levels in in neural tissues of cockroaches exposed to 74 and 740 nM STZ head injection, compared with cockroaches in the control group. All values are mean ± SD. * indicates a significant difference from control; # indicates significant differences from 74 nM STZ injection

Furthermore, we examined the activity of oxidative stress markers in neural tissues and found significantly increased levels of reactive oxygen species (ROS) [F (2,9), = 80.65; P  < 0.0001; Fig.  5 A], malondialdehyde (MDA) [F (2,9), = 514.5; P  < 0.0001; Fig.  5 B], and 2’,7’-dichlorofluorescein (DCF) [F (2,8), = 35.08; P  = 0.0001; Fig.  5 C] in the 74 and 740 nmol STZ treated groups compared with the control group. Antioxidant and detoxification responses were also examined, and we found significantly increased levels of total thiol [F (2,9), = 344.0; P  < 0.0001; Fig.  6 A], and significantly increased activity of Glutathione S-transferase (GST) [F (2,10), = 194.1; P  < 0.0001, Fig.  6 B] in the neural tissues of both 74 and 740 nmol STZ treated cockroaches compared with control. 74 nmol STZ treatment was also significantly different from 740 nmol STZ treatment.

figure 5

Increased oxidative stress, 7 days after head STZ injection in cockroaches. n  = 5. One-way ANOVA with Tukey’s multiple comparisons test indicated a significant increase in ( A ) ROS ( B ) TBARS, and ( C ) DCF levels in neural tissues of cockroaches exposed to 74 and 740 nM STZ head injection, compared with cockroaches in the control group. All values are mean ± SD. * indicates a significant difference from control

figure 6

Increased antioxidant and detoxification activity, 7 days after head STZ injection in cockroaches. n  = 5. One-way ANOVA with Tukey’s multiple comparisons test indicated a significant increase in ( A ) Total Thiol, and ( B ) GST activity in neural tissues of cockroaches exposed to 74 and 740 nM STZ head injection, compared with cockroaches in the control group. All values are mean ± SD. * indicates a significant difference from control. # indicates significant differences from 74 nM STZ injection

Gene expression analyses

We found significantly increased mRNA levels of glucose transporter (GLUT 1) [F (2,14), = 25.20; P  < 0.0001; Fig.  7 A] and phosphoinositide 3-kinase (PI3K) [F (2,14), = 16.54; P  = 0.0002; Fig.  7 B] in the treated groups (74 and 740 nM STZ) compared with control. In the same vein, the mRNA levels of the regulator of ROS generation - Dual oxidase (DUOX) significantly increased in the treated groups compared with the control group [F (2,12), = 50.58; P  < 0.0001; Fig.  8 A], but only the 74 nmol STZ group showed significant increase in the mRNA levels of the detoxification molecule – glutathione S-transferase theta (GST theta) [F (2,14), = 25.20; P  < 0.0001; Fig.  8 B] compared with control. The antioxidant – superoxide dismutase (SOD) [F (2,18), = 20.70; P  < 0.0001; Fig.  8 C] was significantly increased in the 74 and 740 STZ treated groups compared with the control group. Conversely, catalase mRNA levels were significantly reduced in the 74 and 740 nmol STZ treated groups compared with the control group [F (2,13), = 28.54; P  < 0.0001; Fig.  8 D]. Finally, 74 nM head STZ injection induced a significant increase in the mRNA levels of the activators and target genes of the JNK pathway - Early growth response protein 1 [EGR: F (2,16), = 5.075; P  = 0.0196; Fig.  9 A], the TOLL/NF-kB pathway – TOLL 1 [F (2,12), = 6.541; P  = 0.0120; Fig.  9 B], and the UPD3/JAK/STAT pathway - unpaired 3 [UPD3: F (2,10), = 9.345; P  = 0.0051; Fig.  9 C] compared with control.

figure 7

Increased glucose transporter (GLUT 1) activity and phosphoinositide 3-kinase (PI3K) signalling cascade, 7 days after head STZ injection in cockroaches. n  = 9. One-way ANOVA with Tukey’s multiple comparisons test indicated a significant increase in ( A ) GLUT 1, and ( B ) PI3K activity in neural tissues of cockroaches exposed to 74 and 740 nM STZ head injection, compared with cockroaches in the control group. All values are mean ± SD. * indicates a significant difference from control

figure 8

Modulated expression of ROS generation regulator —Dual oxidases (DUOX) and antioxidant/detoxification genes, 7 days after head STZ injection in cockroaches. n  = 9. One-way ANOVA with Tukey’s multiple comparisons test indicated a significant increase in ( A ) DUOX, ( B ) GST Theta, and ( C ) SOD, as well as significant decrease in ( D ) Catalase in neural tissues of cockroaches exposed to 74 and 740 nM STZ head injection, compared with cockroaches in the control group. All values are mean ± SD. * indicates a significant difference from control; # indicates significant differences from 74 nM STZ injection

figure 9

Increased expression of inflammation-associated genes, 7 days after head STZ injection in cockroaches. n  = 9. One-way ANOVA with Tukey’s multiple comparisons test indicated a significant increase in ( A ) EGR ( B ) TOLL 1, and ( C ) UPD3 activity in neural tissues of cockroaches exposed to 74 and 740 nM STZ head injection, compared with cockroaches in the control group. All values are mean ± SD. * indicates a significant difference from control

Insects like Drosophila melanogaster have been exploited to develop insights into several human diseases [ 45 ], and Nauphoeta cinerea has been used to understand several mechanistic responses to neurotoxicant exposure [ 22 , 24 , 25 , 46 , 47 ]. Here, we use N. cinerea – an emerging model of neurotoxicity – to explore neurodegenerative mechanisms, in line with the 3Rs principle to replace, reduce and refine animal models of biomedical research.

Insect survival is rooted in the ability to explore surroundings in the quest for nutrition, mate and escape from predators [ 48 ], but certain xenobiotics interfere with insect eclosion and neuromotor competence [ 18 , 21 , 25 , 49 , 50 ]. For example, organophosphate and pyrethroids inhibit neural acetylcholine breakdown and prolong sodium channel opening respectively, resulting in repeated nerve firing, paralysis and death in insects [ 51 , 52 ]. An in-depth understanding of insect neurotransmitter regulation may then be used to model mammalian neurodegenerative mechanisms. Indeed, dementia is often modelled with intracerebroventricular (ICV) streptozotocin exposure in rats, which impedes brain glucose metabolism and cholinergic transmission [ 53 ]. Streptozotocin exposure reduced N. cinerea survival and explorative abilities, similar to the cognitive and motor deficits that are mapped in ICV STZ treated rodents [ 54 ].

Despite the brain’s critical need for glucose, excessive GLUT 1 activation predisposes to glucose toxicity via dose-dependent mechanisms [ 55 , 56 ], and we found increased expression of GLUT 1, along with increased glucose levels in neural tissues of STZ-treated nymphs. Similarly, increased glucose transporter activity has been recorded in ICV STZ exposed rats, with a consequent suggestion that changes in brain glucose metabolism may predispose to neurodegeneration [ 57 ]. Given that PI3K downstream signaling induces neuronal apoptosis via upregulated AKT and downregulated ERK signaling [ 58 ], and we found increased mRNA expression of PI3K in head STZ injected cockroaches, it is imperative that we further explore PI3K-related neurotoxic mechanisms in the cockroach. Conversely, though hypertriglyceridemia plays a role in insulin resistance that may predispose to neurotoxicity [ 59 ], we found reduced triglyceride levels in neural tissues of STZ head injected cockroaches.

Neurotransmitter regulators are current drug targets for mood and attention deficits in humans, as deficiencies in synaptic and neurotransmitter function have been linked with a spectrum of neurodegenerative diseases [ 60 ]. Consequently, acetylcholinesterase inhibitors impact varying AChE functions, including the modulation of oxidative stress, inflammation, cell apoptosis and adhesion, while monoamine oxidase inhibitors reduce the degradation of biogenic amines like dopamine, noradrenaline and serotonin to enhance mood [ 61 , 62 ]. We recorded increased AChE and MAO activities in head STZ-injected cockroaches. In rodent dementia models, unmitigated AChE release in the synaptic cleft exercabates oxido-inflammatory response and the aggregation of pathological proteins like Aβ peptides [ 62 ], while MAO-A and MAO-B polymorphisms have been linked with several neuropsychiatric diseases [ 63 ]. Although oxidative stress is widely regarded as central to neurodegenerative processes, there are still questions as to whether oxidative stress induces neural degeneration or is a product of dying neurons, and it is still unclear why some nerves are more vulnerable than others [ 64 ]. In this study, head STZ injection increased the expression of the ROS generation regulator – DUOX, with a subsequent increase in levels of oxidative stress markers in neural tissues, including ROS, MDA and DCF. Marked oxidative stress has also been recorded in rat models of cognitive impairment induced by ICV STZ injection [ 26 ]. Indeed, the generation of reactive oxygen and nitrogen species (RONS) from biological processes or exposure to neurotoxicants predispose to the peroxidation of lipid membranes and impaired proteolysis, which may result in protein aggregation and neurodegeneration [ 65 ].

Animal studies have shown promise for antioxidant therapies against neurotoxicity, and we have previously shown N. cinerea’s innate ability to increase antioxidant and detoxification systems when exposed to xenobiotics [ 18 , 21 , 22 , 66 ], which has been replicated here, as we found increased total thiol levels, GST activity and mRNA levels of GST theta and SOD in neural tissues of head STZ-injected cockroaches, although catalase expression was reduced. Cellular homeostasis attempts to balance RONS generation with antioxidant and detoxification mechanism [ 65 ], but this can be overwhelmed by chronic neurotoxicant exposure, and even antioxidant treatments that ameliorate the resulting disease phenotype may not eliminate oxidative stress in rats [ 67 ], which may explain why translational efforts, including large randomized controlled trials in humans, have been unable to repeat the success of antioxidant therapies [ 68 ]. On the other hand, neuroinflammation plays a role in neurotoxicity [ 69 ]. We found increased expression of target genes of the JNK, TOLL and UPD3 inflammation-associated pathways in neural tissues of head STZ-injected cockroaches, like records of proinflammatory cytokine response during neurodegeneration in STZ diabetic rats [ 70 ], thereby lending credence to remedies that target the mechanisms of redox-inflammatory crosstalk in neurodegenerative diseases for therapeutic advantage.

Conclusion and strength of the study

The lobster cockroach can be used to elucidate the mechanisms of chemical-induce neurotoxicity and may potentially serve as a viable model for neurodegeneration, with the possibility of exploiting the cockroach’s neurotransmitter regulators, energy metabolism patterns and redox-inflammatory crosstalk for therapeutic gain. A potentially viable insect model of neurodegeneration also expands the pool of data from which algorithms for new approach methodologies (NAMs) would be built.

Limitation and future perspectives

The neurodegenerative process is multifactorial, and it presents a plethora of phenotypes that require thorough assessments to ensure accurate characterization. The delineation of neuropathological changes in insect models therefore have to mirror mammalian phenotypes. Hence, immunohistochemical analysis of the cockroach brain is a crucial next step in proving the viability of the model.

Data availability

Data used in this study is available upon reasonable request to OBO and OCO.

Lamptey RNL, Chaulagain B, Trivedi R, Gothwal A, Layek B, Singh J. A review of the Common Neurodegenerative disorders: current therapeutic approaches and the potential role of Nanotherapeutics. Int J Mol Sci. 2022;23.

Gilmour GS, Nielsen G, Teodoro T, Yogarajah M, Coebergh JA, Dilley MD, et al. Management of functional neurological disorder. J Neurol. 2020;267:2164–72.

Article   PubMed   PubMed Central   Google Scholar  

Zhang Y, Bebok Z. An examination of mechanisms by which synonymous mutations may alter protein levels, structure and functions. Single Nucleotide Polymorphisms: Hum Variation Coming Revolution Biology Med. 2022;:99–132.

Ogunsuyi OB, Aro OP, Oboh G, Olagoke OC. Curcumin improves the ability of donepezil to ameliorate memory impairment in Drosophila melanogaster: involvement of cholinergic and cnc/Nrf2-redox systems. Drug Chem Toxicol. 2023;46:1035–43.

Article   CAS   PubMed   Google Scholar  

Hirth F. Drosophila melanogaster in the study of human neurodegeneration. CNS Neurol Disord Drug Targets. 2012;9:504–23.

Article   Google Scholar  

Sturtevant AE, Lewis EB. A history of Genetics. Cold Spring Harbour Laboratory; 2001.

Adedara IA, Mohammed KA, Da-Silva OF, Salaudeen FA, Gonçalves FLS, Rosemberg DB, et al. Utility of cockroach as a model organism in the assessment of toxicological impacts of environmental pollutants. Environ Adv. 2022;8:100195.

Article   CAS   PubMed   PubMed Central   Google Scholar  

Cochran D, Organization WH. Cockroaches: their biology, distribution and control. 1999.

Weiler A, Volkenhoff A, Hertenstein H, Schirmeier S. Metabolite transport across the mammalian and insect brain diffusion barriers. Neurobiol Dis. 2017;107:15–31.

Baldwin SA. Mammalian na+-independent sugar transporters and their homologues in other organisms. Cell Physiol Biochem. 1994;4:242–64.

Article   CAS   Google Scholar  

Mattila J, Hietakangas V. Regulation of carbohydrate energy metabolism in Drosophila melanogaster. Genetics. 2017;207:1231–53.

CAS   PubMed   PubMed Central   Google Scholar  

Olagoke OC, Segatto ALA, Afolabi BA, Ardisson-Araujo D, Aschner M, Rocha JBT. RPS6 transcriptional modulation in neural tissues of Nauphoeta cinerea during streptozotocin-associated sugar metabolism impairment. Comp Biochem Physiol B Biochem Mol Biol. 2023;263:110785.

Rulifson EJ, Kim SK, Nusse R. Ablation of insulin-producing neurons in flies: growth and diabetic phenotypes. Science. 2002;296:1118–20.

Bennett RA, Pegg AE. Alkylation of DNA in rat tissues following administration of Streptozotocin. 1981;41 July:2786–91.

Berthiaume JM, Kurdys JG, Muntean DM, Rosca MG. Mitochondrial NAD + /NADH Redox State and Diabetic Cardiomyopathy. Antioxid Redox Signal. 2017;30:375–98.

Article   PubMed   Google Scholar  

Salkovic-Petrisic M, Hoyer S. Central insulin resistance as a trigger for sporadic Alzheimer-like pathology: an experimental approach. J Neural Transmission Supplementa. 2007;:217–33.

Bolzán AD, Bianchi NO, Bianchi MS. Effects of antioxidants on streptozotocin-induced clastogenesis in mammalian and insect cells. Mutat Research/Genetic Toxicol Environ Mutagen. 1998;418:35–42.

Olagoke OC, Afolabi BA, Rocha JBT. Streptozotocin induces brain glucose metabolic changes and alters glucose transporter expression in the lobster cockroach; Nauphoeta cinerea (Blattodea : Blaberidae). Mol Cell Biochem. 2020;:1–35.

Olagoke OC, Segatto ALA, Afolabi BA, Rocha JBT. Streptozotocin activates inflammation-associated signalling and antioxidant response in the lobster cockroach; Nauphoeta cinerea (Blattodea: Blaberidae). Chem Biol Interact. 2021;345:109563.

Mirabeau O, Joly JS. Molecular evolution of peptidergic signaling systems in bilaterians. Proc Natl Acad Sci U S A. 2013;110:E2028–37.

Piccoli BC, Alvim JC, da Silva FD, Nogara PA, Olagoke OC, Aschner M, et al. High level of methylmercury exposure causes persisted toxicity in Nauphoeta cinerea. Environ Sci Pollut Res. 2019. https://doi.org/10.1007/s11356-019-06989-9 .

Afolabi BA, Olagoke OC, Souza DO, Aschner M, Rocha JBT, Segatto ALA. Modified expression of antioxidant genes in lobster cockroach, Nauphoeta cinerea exposed to methylmercury and monosodium glutamate. Chem Biol Interact. 2020;318:108969.

Afolabi BA, Olagoke OC. High concentration of MSG alters antioxidant defence system in lobster cockroach Nauphoeta cinerea (Blattodea: Blaberidae). BMC Res Notes. 2020;13:1–6.

Obafemi BA, Adedara IA, Segatto ALA, Souza DO, da Rocha JBT, Olagoke OC. JNK- and rel-mediated regulation of inflammation and neurotoxicity in Nauphoeta cinerea exposed to Methylmercury and Monosodium Glutamate. J Food Biochem. 2023;2023:1–9.

Olagoke OC, Ogunsuyi OB, Mayokun FE, Rocha JBT, Oboh G. Neurobehavioral, neurotransmitter and redox modifications in Nauphoeta cinerea under mixed heavy metal (silver and mercury) exposure. BMC Research Notes 2024 17:1. 2024;17:1–7.

Sharma M, Gupta YK. Intracerebroventricular injection of streptozotocin in rats produces both oxidative stress in the brain and cognitive impairment. Life Sci. 2001;68:1021–9.

Kramer KJ, Jacobs RM, Speirs RD, Hendmcks LH. Effect of vertebrate hypoglycemic and beta cell cytotoxic agents on insects. Biochem Physiol. 1978;61:95–7.

Google Scholar  

Afolabi BA, Adedara IA, Souza DO, Rocha JBT. Dietary co-exposure to methylmercury and monosodium glutamate disrupts cellular and behavioral responses in the lobster cockroach, Nauphoeta cinerea model. Environ Toxicol Pharmacol. 2018;64:70–7.

Adedara IA, Rosemberg DB, Souza DO, Kamdem JP, Farombi EO, Aschner M, et al. Biochemical and behavioral deficits in the lobster cockroach Nauphoeta cinerea model of methylmercury exposure. Toxicol Res (Camb). 2015;4:442–51.

Silva FDA da, Galiciolli ME, de Irioda A, Oliveira AC, Piccoli CS, de Prestes BC et al. S,. Investigation of the cytotoxicity, genotoxicity and antioxidant prospects of JM-20 on human blood cells: A multi-target compound with potential therapeutic applications. Blood Cells Mol Dis. 2024;106:102827.

Olagoke OC. Modelagem do fenótipo de diabetes em Nauphoeta cinerea usando estreptozotocina: Foco no metabolismo da glicose no cérebro. 2021.

Siddique YH, Naz F, Khan W, Jyoti S, Singh BR, Naqvi AH. Effect of pramipexole alginate nanodispersion (PAND) on the transgenic Drosophila expressing human alpha synuclein in the brain. J Appl Biomed. 2018;16:111–9.

Mcewen CM. Human Plasma Monoamine Oxidase I. PURIFICATION AND IDENTIFICATION. THE IOURNAL OF BIOLWIC ~ L CHEMISTRY. 240.

Pérez-Severiano F, Santamaría A, Pedraza-Chaverri J, Medina-Campos ON, Ríos C, Segovia J. Increased formation of reactive oxygen species, but no changes in glutathione peroxidase activity, in Striata of mice transgenic for the Huntington’s Disease Mutation. Neurochem Res. 2004;29:729–33.

Hayashi I, Morishita Y, Imai K, Nakamura M, Nakachi K, Hayashi T. High-throughput spectrophotometric assay of reactive oxygen species in serum. Mutat Research/Genetic Toxicol Environ Mutagen. 2007;631:55–61.

Oboh G, Ogunsuyi OB, Ojelade MT, Akomolafe SF. Effect of dietary inclusions of bitter Kola seed on geotactic behavior and oxidative stress markers in Drosophila melanogaster. Food Sci Nutr. 2018;6:2177–87.

Ohkawa H, Ohishi N, Yagi K. Assay for lipid peroxides in animal tissues by thiobarbituric acid reaction. Anal Biochem. 1979;95:351–8.

Ogunsuyi OB, Oboh G, Oluokun OO, Ademiluyi AO, Ogunruku OO. Gallic acid protects against neurochemical alterations in transgenic Drosophila model of Alzheimer’s disease. Orient Pharm Exp Med. 2019;20:89–98.

Ellman GL. Tissue sulfhydryl groups. Arch Biochem Biophys. 1959;82:70–7.

Habig WH, Jakoby WB. Assays for differentiation of glutathione S-transferases. Methods Enzymol. 1981;77:398–405.

Ogunsuyi OB, Olagoke OC, Afolabi BA, Oboh G, Ijomone OM, Barbosa NV et al. Dietary inclusions of Solanum vegetables mitigate aluminum-induced redox and inflammation-related neurotoxicity in Drosophila melanogaster model. https://doi.org/101080/1028415X20211933331. 2021;25:2077–91.

Segatto ALA, Diesel JF, Loreto ELS, da Rocha JBT, Segatto ALA, Diesel JF, et al. De novo transcriptome assembly of the lobster cockroach Nauphoeta cinerea (Blaberidae). Genet Mol Biol. 2018;41:713–21.

Ogunsuyi OB, Olagoke OC, Afolabi BA, Oboh G, Ijomone OM, Barbosa NV, et al. Dietary inclusions of Solanum vegetables mitigate aluminum-induced redox and inflammation-related neurotoxicity in Drosophila melanogaster model. Nutr Neurosci. 2022;25:2077–91.

Livak KJ, Schmittgen TD. Analysis of relative gene expression data using real-time quantitative PCR and the 2-∆∆CT method. Methods. 2001;25:402–8.

Oriel C, Lasko P. Recent developments in using Drosophila as a Model for Human Genetic Disease. Int J Mol Sci. 2018;19.

Adedara IA, Mohammed KA, Canzian J, Rosemberg DB, Aschner M, Farombi EO, et al. Nauphoeta cinerea as an emerging model in neurotoxicology. Adv Neurotoxicol. 2023;9:181–96.

Adedara IA, Rosemberg DB, de Souza D, Farombi EO, Aschner M, Souza DO, et al. Neurobehavioral and biochemical changes in Nauphoeta cinerea following dietary exposure to chlorpyrifos. Pestic Biochem Physiol. 2016;130:22–30.

Dingle H, Holyoak M. The evolutionary ecology of movement. Oxford University Press; 2001.

Rand MD, Prince LM, Vorojeikina D, Drosophotoxicology. Elucidating kinetic and dynamic pathways of methylmercury toxicity in a Drosophila model. Front Genet. 2019;10 JUN:444768.

Afolabi BA. Neurotoxicological and molecular aspects of methylmercury and monosodium glutamate exposure in the lobster cockroach, Nauphoeta cinerea. 2019.

Soderlund DM. Molecular mechanisms of pyrethroid insecticide neurotoxicity: recent advances. Arch Toxicol. 2012;86:165–81.

Zhichao W, Zhijiao K, Xueyan S, Xiwu G, Zhichao W, Zhijiao K et al. Research progresses on the metabolic mechanisms of organophosphate insecticides. Chinese Journal of Pesticide Science, 2015, Vol 17, Issue 1, Pages: 1–14. 17:1–14.

Blokland A, Jolles J. Spatial learning deficit and reduced hippocampal ChAT activity in rats after an ICV injection of streptozotocin. Pharmacol Biochem Behav. 1993;44:491–4.

Li D, Huang Y, Cheng B, Su J, Zhou WX, Zhang YX. Streptozotocin induces mild cognitive impairment at appropriate doses in mice as determined by long-term potentiation and the Morris Water Maze. J Alzheimer’s Disease. 2016;54:89–98.

Barros LF, San Martín A, Ruminot I, Sandoval PY, Fernández-Moncada I, Baeza-Lehnert F, et al. Near-critical GLUT1 and neurodegeneration. J Neurosci Res. 2017;95:2267–74.

Tomlinson DR, Gardiner NJ. Glucose neurotoxicity. Nat Rev Neurosci. 2008;9:36–45.

Knezovic A, Loncar A, Homolak J, Smailovic U, Osmanovic Barilar J, Ganoci L, et al. Rat brain glucose transporter-2, insulin receptor and glial expression are acute targets of intracerebroventricular streptozotocin: risk factors for sporadic Alzheimer’s disease? J Neural Transm. 2017;124:695–708.

Cui W, Li W, Han R, Mak S, Zhang H, Hu S, et al. PI3-K/Akt and ERK pathways activated by VEGF play opposite roles in MPP+-induced neuronal apoptosis. Neurochem Int. 2011;59:945–53.

Banks WA, Farr SA, Salameh TS, Niehoff ML, Rhea EM, Morley JE, et al. Triglycerides cross the blood–brain barrier and induce central leptin and insulin receptor resistance. Int J Obes 2018. 2017;42:3.

Huber N, Korhonen S, Hoffmann D, Leskelä S, Rostalski H, Remes AM, et al. Deficient neurotransmitter systems and synaptic function in frontotemporal lobar degeneration—insights into disease mechanisms and current therapeutic approaches. Mol Psychiatry 2021. 2021;27(3):27:1300–9.

Riederer P, Müller T. Use of monoamine oxidase inhibitors in chronic neurodegeneration. Expert Opin Drug Metab Toxicol. 2017;13:233–40.

Walczak-Nowicka ŁJ, Herbet M. Acetylcholinesterase Inhibitors in the Treatment of Neurodegenerative Diseases and the Role of Acetylcholinesterase in their Pathogenesis. International Journal of Molecular Sciences. 2021, Vol 22, Page 9290. 2021;22:9290.

Deftereos SN, Dodou E, Andronis C, Persidis A. Expert Review of Clinical Pharmacology from depression to neurodegeneration and heart failure: re-examining the potential of MAO inhibitors. 2014. https://doi.org/10.1586/ecp.12.29

Gandhi S, Abramov AY. Mechanism of oxidative stress in neurodegeneration. Oxid Med Cell Longev. 2012. https://doi.org/10.1155/2012/428010 .

Dasuri K, Zhang L, Keller JN. Oxidative stress, neurodegeneration, and the balance of protein degradation and protein synthesis. Free Radic Biol Med. 2013;62:170–85.

Olagoke OC, Segatto ALA, Afolabi BA, Rocha JBT. Streptozotocin activates inflammation-associated signalling and antioxidant response in the lobster cockroach; Nauphoeta cinerea (Blattodea: Blaberidae). Chem Biol Interact. 2021;:109563.

Conte V, Uryu K, Fujimoto S, Yao Y, Rokach J, Longhi L, et al. Vitamin E reduces amyloidosis and improves cognitive function in Tg2576 mice following repetitive concussive brain injury. J Neurochem. 2004;90:758–64.

Ienco EC, Logerfo A, Carlesi C, Orsucci D, Ricci G, Mancuso M et al. Oxidative stress treatment for clinical trials in neurodegenerative diseases. J Alzheimers Dis. 2011;24 Suppl 2 SUPPL. 2:111–26.

Glass CK, Saijo K, Winner B, Marchetto MC, Gage FH. Mechanisms underlying inflammation in Neurodegeneration. Cell. 2010;140:918–34.

Ravelli KG, Rosário B, dos A, Vasconcelos AR, Scavone C, Camarini R, Hernandes MS, et al. NADPH oxidase contributes to streptozotocin-induced neurodegeneration. Neuroscience. 2017;358:227–37.

Download references

Acknowledgements

O.B.O is a recipient of the 2022 ICGEB Arthuro-Falaschi (SMART) Postdoctoral Fellowship (S/NGA22-06). The financial support of CAPES, FAPERGS/CNPq and INCT-EN: for Cerebral Diseases, Excitotoxicity, and Neuroprotection is also acknowledged.

This research is funded by the International Center for Genetic Engineering and Biotechnology (ICGEB) and Coordenação de Aperfeiçoamento de Pessoal de Nível Superior (CAPES), Brazil. This work is financially supported by FAPERGS/CNPq 12/2014-PRONEX: nº 16/2551-0000, CAPES/PROEX (n° 23038.004173/2019-93; n° 0493/2019; n° 88882.182125/2018-01; 88882.182123/2018-01), and INCT-EN: for Cerebral Diseases, Excitotoxicity, and Neuroprotection.

Author information

Authors and affiliations.

Departamento de Bioquímica e Biologia Molecular, Programa de Pos-graduacao em Bioquimica Toxicologica, Centro de Ciências Naturais e Exatas (CCNE), Universidade Federal de Santa Maria, Santa Maria, RS, 97105-900, Brazil

Opeyemi B. Ogunsuyi, Nilda V. Barbosa & João B.T. Rocha

Department of Biomedical Technology, Federal University of Technology, P.M.B. 704, Akure, Nigeria

Opeyemi B. Ogunsuyi & Mayokun E. Famutimi

Drosophila Research Lab, Functional Foods and Nutraceuticals Unit, Federal University of Technology, P.M.B. 704, Akure, Nigeria

Opeyemi B. Ogunsuyi, Mayokun E. Famutimi, Damilola M. Olatunde & Ganiyu Oboh

Department of Medicine, Division of Gastroenterology, Beth Israel Deaconess Medical Center, Harvard Medical School, Boston, MA, USA

Olawande C. Olagoke

Department of Medicine, Division of Translational Research and Technology Innovation, Beth Israel Deaconess Medical Center, Harvard Medical School, Boston, MA, USA

Department of Physiology, Kampala International University, Ishaka-Bushenyi, Uganda

Department of Biochemistry, Federal University of Technology, P.M.B. 704, Akure, Nigeria

Damilola M. Olatunde & Ganiyu Oboh

Departamento de Bioquímica, Instituto de Ciências Básicas da Saúde, Universidade Federal do Rio Grande do Sul, Rua Ramiro Barcelos 2600-Anexo, Porto Alegre, RS, 90035-003, Brazil

Diogo O. G. Souza & João B.T. Rocha

You can also search for this author in PubMed   Google Scholar

Contributions

O.B.O conducted lab experiments and analysed the results, O.C.O interpreted the results and wrote the manuscript, M.E.F conducted lab experiments, D.M.O conducted lab experiments, D.O.S edited the manuscript, G.O edited the manuscript, N.V.B edited the manuscript, J.B.T edited the manuscript. All authors reviewed the manuscript.

Corresponding author

Correspondence to Olawande C. Olagoke .

Ethics declarations

Ethics approval and consent to participate.

Not applicable.

Consent to publish

Competing interests.

The authors declare no competing interests.

Additional information

Publisher’s note.

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/ .

Reprints and permissions

About this article

Cite this article.

Ogunsuyi, O.B., Olagoke, O.C., Famutimi, M.E. et al. Neural acetylcholinesterase and monoamine oxidase deregulation during streptozotocin-induced behavioral, metabolic and redox modification in Nauphoeta cinerea . BMC Neurosci 25 , 42 (2024). https://doi.org/10.1186/s12868-024-00890-z

Download citation

Received : 07 April 2024

Accepted : 23 August 2024

Published : 29 August 2024

DOI : https://doi.org/10.1186/s12868-024-00890-z

Share this article

Anyone you share the following link with will be able to read this content:

Sorry, a shareable link is not currently available for this article.

Provided by the Springer Nature SharedIt content-sharing initiative

  • Redox-inflammation crosstalk
  • Neurodegeneration
  • Oxidative damage
  • Neurotoxicity

BMC Neuroscience

ISSN: 1471-2202

redox reaction and catalase activity experiment

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • View all journals
  • Explore content
  • About the journal
  • Publish with us
  • Sign up for alerts
  • Published: 02 September 2024

Gold nugget formation from earthquake-induced piezoelectricity in quartz

  • Christopher R. Voisey   ORCID: orcid.org/0009-0009-2693-2301 1 ,
  • Nicholas J. R. Hunter 1 , 2 ,
  • Andrew G. Tomkins 1 ,
  • Joël Brugger   ORCID: orcid.org/0000-0003-1510-5764 1 ,
  • Weihua Liu   ORCID: orcid.org/0000-0002-2091-7137 3 ,
  • Yang Liu   ORCID: orcid.org/0000-0002-0750-7571 4 &
  • Vladimir Luzin   ORCID: orcid.org/0000-0003-2635-6921 5  

Nature Geoscience ( 2024 ) Cite this article

255 Altmetric

Metrics details

  • Economic geology
  • Geochemistry

Gold nuggets occur predominantly in quartz veins, and the current paradigm posits that gold precipitates from dilute (<1 mg kg −1 gold), hot, water ± carbon dioxide-rich fluids owing to changes in temperature, pressure and/or fluid chemistry. However, the widespread occurrence of large gold nuggets is at odds with the dilute nature of these fluids and the chemical inertness of quartz. Quartz is the only abundant piezoelectric mineral on Earth, and the cyclical nature of earthquake activity that drives orogenic gold deposit formation means that quartz crystals in veins will experience thousands of episodes of deviatoric stress. Here we use quartz deformation experiments and piezoelectric modelling to investigate whether piezoelectric discharge from quartz can explain the ubiquitous gold–quartz association and the formation of gold nuggets. We find that stress on quartz crystals can generate enough voltage to electrochemically deposit aqueous gold from solution as well as accumulate gold nanoparticles. Nucleation of gold via piezo-driven reactions is rate-limiting because quartz is an insulator; however, since gold is a conductor, our results show that existing gold grains are the focus of ongoing growth. We suggest this mechanism can help explain the creation of large nuggets and the commonly observed highly interconnected gold networks within quartz vein fractures.

This is a preview of subscription content, access via your institution

Access options

Access Nature and 54 other Nature Portfolio journals

Get Nature+, our best-value online-access subscription

24,99 € / 30 days

cancel any time

Subscribe to this journal

Receive 12 print issues and online access

251,40 € per year

only 20,95 € per issue

Buy this article

  • Purchase on SpringerLink
  • Instant access to full article PDF

Prices may be subject to local taxes which are calculated during checkout

redox reaction and catalase activity experiment

Similar content being viewed by others

redox reaction and catalase activity experiment

Cathodoluminescence as a tracing technique for quartz precipitation in low velocity shear experiments

redox reaction and catalase activity experiment

Transport and coarsening of gold nanoparticles in an orogenic deposit by dissolution–reprecipitation and Ostwald ripening

redox reaction and catalase activity experiment

Formation of orogenic gold deposits by progressive movement of a fault-fracture mesh through the upper crustal brittle-ductile transition zone

Data availability.

Neutron diffraction measurements of quartz samples and associated data used to model the piezoelectric properties are available via figshare at https://doi.org/10.6084/m9.figshare.26315281 (ref. 50 ). All other data supporting the findings of this study (sample images and geochemical models) are available within the article and its extended data files.

Code availability

The code used for piezoelectric potential tensorial modelling can be found in ref. 20 .

Phillips, N. in Modern Approaches in Solid Earth Sciences (eds Dilek, Y., Pirajno, F. & Windley, B.) 7–15 (Springer, 2022).

Sibson, R. H., Moore, J. M. & Rankin, A. H. Seismic pumping—a hydrothermal fluid transport mechanism. J. Geol. Soc. 131 , 653–659 (1975).

Article   Google Scholar  

Cox, S. F. Faulting processes at high fluid pressures: an example of fault valve behavior from the Wattle Gully Fault, Victoria, Australia. J. Geophys. Res. Solid Earth 100 , 12841–12859 (1995).

Groves, D. I. The crustal continuum model for late-archaean lode-gold deposits of the Yilgarn Block, Western Australia. Miner. Deposita 28 , 366–374 (1993).

Tomkins, A. G. & Grundy, C. Upper temperature limits of orogenic gold deposit formation: constraints from the granulite-hosted Griffin’s Find Deposit, Yilgarn Craton. Econ. Geol. 104 , 669–685 (2009).

Article   CAS   Google Scholar  

Cox, S. F. & Etheridge, M. A. Crack–seal fibre growth mechanisms and their significance in the development of oriented layer silicate microstructures. Tectonophysics 92 , 147–170 (1983).

Ramsay, J. G. The crack–seal mechanism of rock deformation. Nature 284 , 135–139 (1980).

Renard, F., Gratier, J. P. & Jamtveit, B. Kinetics of crack-sealing, intergranular pressure solution, and compaction around active faults. J. Struct. Geol. 22 , 1395–1407 (2000).

Williams-Jones, A. E., Bowell, R. J. & Migdisov, A. A. Gold in solution. Elements 5 , 281–287 (2009).

Voisey, C. R. et al. Aseismic refinement of orogenic gold systems. Econ. Geol. 115 , 33–50 (2020).

Petrella, L. et al. Nanoparticle suspensions from carbon-rich fluid make high-grade gold deposits. Nat. Commun. 13 , 3795 (2022).

Williams, R. T. Coseismic boiling cannot seal faults: implications for the seismic cycle. Geology 47 , 461–464 (2019).

Curie, J. & Curie, P. Développement par compression de l’électricité polaire dans les Cristaux hémièdres à faces inclinées. Bull. Soc. Minéral. France 3 , 90–93 (1880).

Bishop, J. R. Piezoelectric effects in quartz-rich rocks. Tectonophysics 77 , 297–321 (1981).

Finkelstein, D. & Powell, J. Earthquake lightning. Nature 228 , 759–760 (1970).

Wang, J. H. Piezoelectricity as a mechanism on generation of electromagnetic precursors before earthquakes. Geophys. J. Int. 224 , 682–700 (2020).

Ghomshei, M. M. & Templeton, T. L. Piezoelectric and a -axes fabric along a quartz vein. Phys. Earth Planet. Inter. 55 , 374–386 (1989).

Parkhomenko, E. I. Electrification Phenomena in Rocks (Pleunum Press, 1971).

Murthy, Y. S. & Bhimasankaram, V. L. Experimental results of the piezoelectric activity of Quartzose Rocks. in SEG Technical Program Expanded Abstracts (Society of Exploration Geophysicists, 1985); https://doi.org/10.1190/1.1892893

Mainprice, D., Bachmann, F., Hielscher, R., Schaeben, H. & Lloyd, G. E. Calculating anisotropic piezoelectric properties from texture data using the MTEX open-source package. Geol. Soc. Lond. Spec. Publ. 409 , 223–249 (2014).

Starr, M. B. & Wang, X. Coupling of piezoelectric effect with electrochemical processes. Nano Energy 14 , 296–311 (2015).

Wang, Z. L. Piezopotential gated nanowire devices: piezotronics and piezo-phototronics. Nano Today 5 , 540–552 (2010).

Wang, Z. L. Progress in piezotronics and piezo-phototronics. Adv. Mater. 24 , 4632–4646 (2012).

Zhang, Y., Liu, Y. & Wang, Z. L. Fundamental theory of piezotronics. Adv. Mater. 23 , 3004–3013 (2011).

Gerischer, H. & Ekardt, W. Fermi levels in electrolytes and the absolute scale of redox potentials. Appl. Phys. Lett. 43 , 393–395 (1983).

Leblanc, S. E. & Fogler, H. S. The role of conduction/valence bands and redox potential in accelerated mineral dissolution. AlChE J. 32 , 1702–1709 (1986).

Hong, K.-S., Xu, H., Konishi, H. & Li, X. Direct water splitting through vibrating piezoelectric microfibers in water. J. Phys. Chem. Lett. 1 , 997–1002 (2010).

Hong, K.-S., Xu, H., Konishi, H. & Li, X. Piezoelectrochemical effect: a new mechanism for azo dye decolorization in aqueous solution through vibrating piezoelectric microfibers. J. Phys. Chem. C 116 , 13045–13051 (2012).

Starr, M. B., Shi, J. & Wang, X. Piezopotential-driven redox reactions at the surface of piezoelectric materials. Angew. Chem. Int. Ed. 51 , 5962–5966 (2012).

Cox, S. F. & Ruming, K. The St Ives Mesothermal Gold System, Western Australia—a case of golden aftershocks? J. Struct. Geol. 26 , 1109–1125 (2004).

Tosi, P., Sbarra, P. & De Rubeis, V. Earthquake sound perception. Geophys. Res. Lett. 39 , 24 (2012).

Pokrovski, G. S., Akinfiev, N. N., Borisova, A. Y., Zotov, A. V. & Kouzmanov, K. Gold speciation and transport in geological fluids: Insights from experiments and physical–chemical modelling. Geol. Soc. Lond. Spec. Publ. 402 , 9–70 (2014).

Bard, A. J., Parsons, R. & Jordan, J. Standard Potentials in Aqueous Solution (Routledge, 2017).

Maddox, L. M., Bancroft, G. M., Scaini, M. J. & Lorimer, J. W. Invisible gold; comparison of au deposition on pyrite and arsenopyrite. Am. Mineral. 83 , 1240–1245 (1998).

Liu, D., Zhou, W., Song, X. & Qiu, Z. Fractal simulation of flocculation processes using a diffusion-limited aggregation model. Fractal Fract. 1 , 12 (2017).

Witten, T. A. & Sander, L. M. Diffusion-limited aggregation, a kinetic critical phenomenon. Phys. Rev. Lett. 47 , 1400–1403 (1981).

Saunders, J. A. & Schoenly, P. A. Boiling, colloid nucleation and aggregation, and the genesis of bonanza Au–Ag ores of the Sleeper Deposit, Nevada. Miner. Deposita 30 , 199–210 (1995).

Monecke, T. et al. Natural growth of gold dendrites within silica gels. Geology 51 , 189–192 (2023).

Polte, J. Fundamental growth principles of colloidal metal nanoparticles—a new perspective. CrystEngComm 17 , 6809–6830 (2015).

Roustom, B. E., Fóti, G. & Comninellis, C. Preparation of gold nanoparticles by heat treatment of sputter deposited gold on boron-doped diamond film electrode. Electrochem. Commun. 7 , 398–405 (2005).

Robert, F., Boullier, A.-M. & Firdaous, K. Gold–quartz veins in metamorphic terranes and their bearing on the role of fluids in faulting. J. Geophys. Res. Solid Earth 100 , 12861–12879 (1995).

Goldfarb, R. J., Groves, D. I. & Gardoll, S. Orogenic gold and geologic time: a global synthesis. Ore Geol. Rev. 18 , 1–75 (2001).

Micklethwaite, S., Sheldon, H. A. & Baker, T. Active fault and shear processes and their implications for mineral deposit formation and discovery. J. Struct. Geol. 32 , 151–165 (2010).

Starr, M. B. & Wang, X. Fundamental analysis of piezocatalysis process on the surfaces of strained piezoelectric materials. Sci. Rep. 3 , 2160 (2013).

Trepmann, C. A. & Stöckhert, B. Short-wavelength undulatory extinction in quartz recording coseismic deformation in the middle crust—an experimental study. Solid Earth 4 , 263–276 (2013).

Hunter, N. J. R. et al. Deformation mechanisms in orogenic gold systems during aseismic periods: microstructural evidence from the central Victorian gold deposits, Southeast Australia. Econ. Geol. 116 , 1849–1864 (2021).

Bunge, H.-J. Texture Analysis in Materials Science: Mathematical Methods (Elsevier, 1982).

Goldfarb, R. J. & Groves, D. I. Orogenic gold: common or evolving fluid and metal sources through time. Lithos 233 , 2–26 (2015).

Liu, W. et al. Colloidal gold in sulphur and citrate-bearing hydrothermal fluids: an experimental study. Ore Geol. Rev. 114 , 103142 (2019).

Voisey, C. Piezo_properties.xlsx. figshare https://doi.org/10.6084/m9.figshare.26315281 (2024).

Download references

Acknowledgements

This study is supported by Australian Research Council (LP200200897) and MRIWA project M10412 awarded to A.G.T., J.B. and W.L. We acknowledge the use of instruments and scientific and technical assistance at the Monash Centre for Electron Microscopy (MCEM), Monash University, the Victorian Node of Microscopy Australia. This research used equipment funded by Australian Research Council grant: Thermo Fisher Scientific Helios 5 UX FIB-SEM ARC Funding (LE200100132). We thank Agnico Eagles Mines Limited and the staff at Fosterville Gold Mine for providing samples and site access. We thank Y. Xing and D. Willis for assistance and discussions throughout the project.

Author information

Authors and affiliations.

School of Earth, Atmosphere and Environment, Monash University, Melbourne, Victoria, Australia

Christopher R. Voisey, Nicholas J. R. Hunter, Andrew G. Tomkins & Joël Brugger

Centre for Health Systems Development, Australian Institute of Primary Care and Ageing, La Trobe University, Bundoora Campus, Melbourne, Victoria, Australia

Nicholas J. R. Hunter

CSIRO Mineral Resources, Clayton, Victoria, Australia

Monash Centre for Electron Microscopy, Monash University, Melbourne, Victoria, Australia

Australian Centre for Neutron Scattering, ANSTO, Sydney, New South Wales, Australia

Vladimir Luzin

You can also search for this author in PubMed   Google Scholar

Contributions

C.R.V. conceptualized the project, designed and conducted piezoelectric laboratory experiments and was lead writer of the paper. N.J.R.H. processed and interpreted neutron diffraction data, constructed relevant figures and helped write the paper. A.G.T. helped conceptualize the project and interpretation. J.B. helped design the aqueous experiments and solutions and constructed geochemical models. W.L. designed and created the nanoparticle suspensions. Y.L. conducted SEM and energy-dispersive spectroscopy. V.L. conducted the neutron diffraction experiments. All authors reviewed the paper before submission.

Corresponding author

Correspondence to Christopher R. Voisey .

Ethics declarations

Competing interests.

The authors declare no competing interests.

Peer review

Peer review information.

Nature Geoscience thanks David Groves, Mark Hannington, Randolph Williams and Yanhao Yu for their contribution to the peer review of this work. Primary Handling Editor: Alison Hunt, in collaboration with the Nature Geoscience team.

Additional information

Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Extended data

Extended data fig. 1 the crystallography of quartz and related piezoelectric effects..

The crystal planes of quartz and the piezoelectric response when distorted. ( a ) Crystallography of left- and right- handed quartz. The basal {c}, prismatic {m} and rhombohedral {r, z} planes are indicated. Minor planes, such as bipyramidal and acute rhombohedral, are not shown. ( b ) Quartz crystal viewed parallel to the c-axis. Here, the first- {m} and second- order {a} prismatic planes can be easily distinguished. ( c ) The effect of an applied mechanical stress (parallel to X) on the quartz atomic framework (top left and right). As the framework is distorted (top right), a piezoelectric potential is generated. When a quartz crystal is viewed parallel to the c-axis (bottom left and right) the distribution of positive (red) and negative (blue) piezoelectric charge can be recognised. ( d ) Note that only the {a} crystal planes are piezoelectric in quartz.

Extended Data Fig. 2 Geological map of the Victorian goldfields, Australia.

Major gold deposits in the Victorian goldfields, including the Fosterville deposit where samples for this study were sourced, from Voisey et al. (2020). ( a ) Schematic of Australia with the state of Victoria highlighted in grey. Position of ( b ) is indicated. ( b ) Inset map of Victoria and part of New South Wales showing the locations of the Lachlan orogen and the Delamerian orogen. Position of ( c ) is indicated. ( c ) Simplified geologic map of central Victoria, modified from Phillips et al. (2012). Infilled circles show major gold fields.

Extended Data Fig. 3 Schematic of the apparatus used in all experiments.

Deformation apparatus used in our experiments. The 2 x 1 x 0.5 cm quartz slab(s) are placed within an 8 x 8 x 3 cm sample chamber and submerged in 75 ml of gold-bearing solution. The perimeter of the sample chamber is sealed with silicon and the bottom plate has a trough to keep the sample chamber in place. Pressure is applied between the bottom two plates to prevent vertical bouncing during experimental oscillations. The quartz is then deformed by the actuator impact head for 1 hour at room temperature.

Extended Data Fig. 4 Gold solubility in our experiments vs. typical orogenic systems.

Eh vs. pH diagrams showing the potentials required to reduce aqueous gold. ( a ) Gold present as AuCl 4 − in our room-temperature experiments and ( b ) shown as the Au(HS) 2 - ± Au(HS)(aq) complexes typical in orogenic gold fluids, into metallic gold, as a function of gold in solution. The diagrams of iron also shown for comparison, where ( c ) and ( d ) correspond to ( a ) and ( b ), respectively. Abbreviations: Hem – hematite, Mgn – magnetite, Po- pyrrhotite, Py – pyrite. QMF in ( d ) shows the pH corresponding to quartz-muscovite-K-feldspar for activities of K + of 0.1 to 0.01.

Extended Data Fig. 5 Control results from uncoated quartz experiments.

Imagery of the quartz crystal control slabs from our uncoated experiments. Samples were submerged in their respective solutions, but not deformed. ( a ) BSE image of bare quartz gold chloride (AuCl4) experiment. ( b ) and ( c ) are BSE and SE images, respectively, of the square area outlined in ( a ). ( d ) BSE image of bare quartz gold nanoparticle (AuNP) experiment. ( e ) and ( f ) are BSE and SE images, respectively, of the square area outlined in ( d ). BSE: Backscattered electron. SE: Secondary electron.

Extended Data Fig. 6 Results from Ir-coated quartz with gold chloride (AuCl 4 ) experiment.

Imagery of the Ir-coated quartz crystal slab after deformation within AuCl 4 solution. ( a ) BSE image of the quartz surface exhibiting distribution of gold particles deposits from AuCl 4 solution. Linear arrays, or ‘branches’, of gold particles can be seen. ( b ) and ( c ) are BSE and SE images, respectively, of the square area outlined in ( a ). Coupling of gold particles is evident as well as a pseudo-hexagonal Au nanocrystal. ( d ) EDS image of the square area in ( a ) highlighting the chemistry of sample area. BSE: Backscattered electron. SE: Secondary electron. EDS: Energy dispersive spectroscopy.

Extended Data Fig. 7 Control results from from Ir-coated quartz with gold chloride (AuCl 4 ) experiment.

Imagery of the Ir-coated quartz crystal control slab for the AuCl 4 solution experiment. Sample was submerged but not deformed. ( a ) EDS image of the quartz control sample surface. ( b ) BSE image of the quartz control sample surface. Inset shown at higher magnification. ( c ) EDS spectra of area in ( a ). BSE: Backscattered electron. EDS: Energy dispersive spectroscopy.

Extended Data Fig. 8 Control results from natural auriferous quartz experiments.

Imagery of the natural gold-bearing quartz control slabs from our growth experiments. Samples were submerged in their respective solutions, but not deformed. ( a ) BSE image of natural auriferous quartz gold chloride (AuCl4) experiment. Gold grain within quartz. ( b ) and ( c ) are BSE and SE images, respectively, of the square area outlined in ( a ). ( d ) BSE image of natural auriferous quartz gold nanoparticle (AuNP) experiment. Gold grain within quartz. ( e ) and ( f ) are BSE and SE images, respectively, of the square area outlined in ( d ). In both samples, particles seen on the gold grain surface are pieces of quartz. BSE: Backscattered electron. SE: Secondary electron.

Extended Data Fig. 9 Results from Ir-coated quartz gold nanoparticle (AuNP) experiment.

Imagery of the Ir-coated quartz crystal slab after deformation within AuNP solution. ( a ) BSE image of the quartz surface exhibiting distribution of gold particles deposits from AuNP solution. Large clusters of AuNPs can be seen. ( b ) and ( c ) are BSE and SE images, respectively, of the square area outlined in ( a ). ( d ) EDS image of the square area in ( a ) highlighting the chemistry of sample area. BSE: Backscattered electron. SE: Secondary electron. EDS: Energy dispersive spectroscopy.

Extended Data Fig. 10 Control results from Ir-coated quartz with gold nanoparticle (AuNP) experiment.

Imagery of the Ir-coated quartz crystal control slab for the AuNP solution experiment. Sample was submerged but not deformed. ( a ) EDS image of the quartz control sample surface. ( b ) BSE image of the quartz control sample surface. Inset shown at higher magnification. ( c ) EDS spectra of area in ( a ). BSE: Backscattered electron. EDS: Energy dispersive spectroscopy.

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Cite this article.

Voisey, C.R., Hunter, N.J.R., Tomkins, A.G. et al. Gold nugget formation from earthquake-induced piezoelectricity in quartz. Nat. Geosci. (2024). https://doi.org/10.1038/s41561-024-01514-1

Download citation

Received : 12 September 2023

Accepted : 22 July 2024

Published : 02 September 2024

DOI : https://doi.org/10.1038/s41561-024-01514-1

Share this article

Anyone you share the following link with will be able to read this content:

Sorry, a shareable link is not currently available for this article.

Provided by the Springer Nature SharedIt content-sharing initiative

Quick links

  • Explore articles by subject
  • Guide to authors
  • Editorial policies

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

redox reaction and catalase activity experiment

IMAGES

  1. Redox Reaction Experiment

    redox reaction and catalase activity experiment

  2. Catalase Activity Assay

    redox reaction and catalase activity experiment

  3. REDOX REACTIONS AND ELECTRODE PROCESSES

    redox reaction and catalase activity experiment

  4. Proposed mechanism for catalase activity with metal-centered redox

    redox reaction and catalase activity experiment

  5. Experiment 10 oxidation-reduction reactions 519

    redox reaction and catalase activity experiment

  6. Catalase enzyme activity experiment

    redox reaction and catalase activity experiment

VIDEO

  1. oxidation and reduction reaction

  2. Redox Reaction

  3. CATALASE- REDOX METALLOENZYME

  4. Reacciones Redox Actividad

  5. Catalase Lab (day 1)

  6. The Catalase Reaction

COMMENTS

  1. Investigating an enzyme-controlled reaction: catalase and hydrogen

    Hydrogen peroxide is harmful and must be removed as soon as it is produced in the cell. Cells make the enzyme catalase to remove hydrogen peroxide. This investigation looks at the rate of oxygen production by the catalase in pureed potato as the concentration of hydrogen peroxide varies. The oxygen produced in 30 seconds is collected over water.

  2. LAB BIO400 Experiment 3; Redox- Reaction Catalase Activity

    This experiment was conducted to study the redox reaction, catalase activity in living tissue and the effect of temperature on catalase activity. For the redox reaction, three small slices of potato and apple were cut and put on three different petri-dish. The first slice was exposed to air, the second slice was put in water while the third ...

  3. 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.

  4. 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).

  5. PDF 5.310 (F19) Catalase Kinetics Lab Manual

    A Study of the Kinetics of the Enzyme Catalase and its Reaction . With H. 2 O 2. 2 A Further Study on Protein Assay Quantitation of Catalase 3 I. OVERVIEW OF THE EXPERIMENT In this experiment, the student will investigate the enzyme activity of catalase by . studying the decomposition of H 2 O. 2 to form water and oxygen. Using an oxygen based

  6. Direct Measurement of Catalase Activity in Living Cells and Tissue

    Figure 3. Direct measurement of Catalase activity in freshly excised tissue specimens in real-time : (a) Oxygen release profiles upon stimulation with 200μM H 2 O 2. Young mouse (~4 weeks old) liver tissues were homogenized to yield protein lysates. Protein concentration was measured by standard laboratory assays.

  7. Enzyme Action: Testing Catalase Activity

    Objectives. In this experiment, you will. Use a Gas Pressure Sensor to measure the production of oxygen gas as hydrogen peroxide is destroyed by the enzyme catalase or peroxidase at various enzyme concentrations. Measure and compare the initial rates of reaction for this enzyme when different concentrations of enzyme react with H 2 O 2. Measure ...

  8. A comprehensive evaluation of catalase-like activity of different

    Assessment of the catalase-like activity of redox-active therapeutics. Experimental set-up for the determination of catalase-like activity (A); The determination of the k obs from the dependence of initial rates, ν 0 on the concentration of either catalase enzyme (B) or MnTnBuOE-2-PyP 5+ (C). Experiments were carried out at (25±1)°C in 0.05 ...

  9. Investigating activation energies

    Calculating activation energies. For the catalase-catalysed reaction, the relationship between reaction rate and concentration of hydrogen peroxide can be represented as rate = k[H 2 O 2] x [catalase] y, where k is the rate constant, the exponent, x, is the order of the reaction with respect to hydrogen peroxide, and the exponent, y, is the order of the reaction with respect to catalase.

  10. A Simple Assay for Measuring Catalase Activity: A Visual Approach

    In this study, an assay that combines the ease and simplicity of the qualitative approach for measuring catalase activity was developed. The assay reagents comprised only hydrogen peroxide and ...

  11. Catalase, a remarkable enzyme: targeting the oldest antioxidant enzyme

    The catalase activity ... currently employed in immunoblotting experiments, is one example of catalase-peroxidase. Manganese catalases ... it reacts rapidly with a second molecule of H 2 O 2 to generate H 2 O and O 2 in a two-electron redox process. This second reaction is particularly efficient in some catalases compared to other heme proteins ...

  12. The Molecular Mechanism of the Catalase Reaction

    Catalases are ubiquitous enzymes that prevent cell oxidative damage by degrading hydrogen peroxide to water and oxygen (2H2O2 → 2 H2O + O2) with high efficiency. The enzyme is first oxidized to a high-valent iron intermediate, known as Compound I (Cpd I) which, in contrast to other hydroperoxidases, is reduced back to the resting state by further reacting with H2O2. By means of hybrid QM/MM ...

  13. Mechanisms of oxidant generation by catalase

    Further insight into potential mechanisms mediating ROS generation was provided by inhibitor studies using mouse and human catalases. As previously mentioned, it is well established that catalase possesses an activity responsible for degrading hydrogen peroxide via the reaction: 2H 2 O 2 → 2H 2 O + O 2. 19-21 In addition, in the presence of low con-centrations of hydrogen peroxide, less ...

  14. Enzyme Action: Catalase Lab

    Design a series of experiments to investigate how differing concentrations of the substrate hydrogen peroxide might affect the rate of enzyme activity. Design an experiment to determine the effect of boiling the catalase on the reaction rate. Explain how environmental factors affect the rate of enzyme-catalyzed reactions.

  15. Therapeutic potentials of catalase: Mechanisms, applications, and

    This article explores the therapeutic potentials of catalase, focusing on its mechanisms, applications, and future perspectives. Catalase, an enzyme present in nearly all living organisms exposed to oxygen, plays a crucial role in protecting cells from the harmful effects of ROS. Its significance in cellular physiology extends beyond mere ...

  16. TASK 1: REDOX REACTION AND CATALASE ACTIVITY

    TASK 1: REDOX REACTION AND CATALASE ACTIVITY by shahirah adnan on Prezi. Blog. July 25, 2024. Sales pitch presentation: creating impact with Prezi. July 22, 2024. Make every lesson count with these student engagement strategies. July 18, 2024. Product presentations: defining them and creating your own.

  17. Designed EXP- Redox Reaction AND Catalase Activity GP 1

    TASK 2: Comparing catalase activity in living tissues. 2ml of hydrogen peroxide was placed into 5 clean tests tub. Then for each test tube was containing different thing which is the first test tube contain small piece of potato, second contain small piece of potato, and third contain chicken meat.

  18. PDF Experiment 12 Redox Reactions

    Experiment 12 Redox Reactions OUTCOMES After completing this experiment, the student should be able to: develop an activity series for different elements and ions. write balanced redox equations. describe the effect of pH on the degree of oxidation or reduction that takes place in a reaction. DISCUSSION

  19. PDF Chem 112 OXIDATION-REDUCTION EXPERIMENT

    Chem 112 OXIDATION-REDUCTION EXPERIMENT. INTRODUCTION . An oxidation-reduction (redox) reaction involves the movement of electrons from one reactant to another. Many reactions that you have already studied are redox reactions; these include single replacement, combustion, and combination. Oxidation is the loss of electrons.

  20. Effector CLas0185 targets methionine sulphoxide reductase B1 of Citrus

    The experiment contained four independent biological replicates, and three technical repeats were performed. (c-e) Detection of ascorbate peroxidase (APX) activity. APX activity was detected in citrus plants expressing CLas0185 and CsMsrB1, as well as in rootlets co-expressing CLas0185 and CsMsrB1. Wild type (WT) served as the control for ...

  21. High‐Entropy Catalysis Accelerating Stepwise Sulfur Redox Reactions for

    Catalysis is crucial to improve redox kinetics in lithium-sulfur (Li-S) batteries. However, conventional catalysts that consist of a single metal element are incapable of accelerating stepwise sulfur redox reactions which involve 16-electron transfer and multiple Li 2 S n (n = 2-8) intermediate species. To enable fast kinetics of Li-S batteries, it is proposed to use high-entropy alloy ...

  22. Revealing the degradation pathways of layered Li-rich oxide cathodes

    These oxidation-state maps allow us to quantitatively evaluate or track redox reactions that occurred in the first cycle (as shown in Fig. 3b-e) and provide visible evidence that oxygen release ...

  23. Conformational control over proton-coupled electron transfer in

    A time-resolved iron-specific x-ray absorption experiment yields no evidence for an Fe 2+ → Fe 3+ transition during Q A − → Q B electron transfer in the photosynthetic reaction center ...

  24. Role of Catalase in Oxidative Stress- and Age-Associated Degenerative

    Conclusion. Catalase is one of the crucial antioxidant enzymes which plays an important role by breaking down hydrogen peroxide and maintaining the cellular redox homeostasis. Diabetes, Alzheimer's disease, Parkinson's disease, etc. are currently becoming common diseases.

  25. Diagnostics

    Hepatic fibrosis with various origins can be estimated non-invasively by using certain biomarkers and imaging-based measurements. The aim of our study was to examine redox homeostasis biomarkers and liver stiffness measurements for the assessment of significant liver fibrosis in different etiologies of chronic liver diseases. A cohort study consisting of 88 chronic liver disease patients of ...

  26. Annexins: A family of calcium binding proteins with variety of roles in

    The research of plant annexins and its involvement in growth, development, and responses to environmental stresses has advanced significantly over the past 20 years in response to the global increase of harsh climates that cause natural disasters (Williams et al., 2020; Saad et al., 2020).Many transgenic experiments have demonstrated a beneficial effect of annexin gene expression on plant ...

  27. Neural acetylcholinesterase and monoamine oxidase deregulation during

    The AChE activity was thereafter calculated and expressed as mmolAcSch/h/mg protein . Monoamine oxidase (MAO) activity was determined as previously reported [32, 33]. The reaction consisted (final concentration) of 72 mM potassium phosphate buffer (pH 7.4), 0. 5 mM benzylamine, 50 µl of tissue homogenate, and 50 µl of distilled water.

  28. Gold nugget formation from earthquake-induced piezoelectricity in

    The elastic rattling of quartz crystals within veins during seismic activity leads to oscillations of the oxidation-reduction reaction fronts on either side of quartz crystal surfaces (Fig. 2).