Lock-and-key model

strong>Lock-and-key model n., [lɑk ænd ki ˈmɑdl̩] Definition: a model for enzyme-substrate interaction

Table of Contents

Lock-and-key model Definition

Lock-and-key model is a model for enzyme-substrate interaction suggesting that the enzyme and the substrate possess specific complementary geometric shapes that fit exactly into one another. In this model, enzymes are depicted as highly specific. They must bind to specific substrates before they catalyze chemical reactions . The term is a pivotal concept in enzymology to elucidate the intricate interaction between enzymes and substrates at the molecular level. In the lock-and-key model, the enzyme-substrate interaction suggests that the enzyme and the substrate possess specific complementary geometric shapes that fit exactly into one another. Like a key  into a  lock , only the correct size and shape of the substrate ( the key ) would fit into the  active site  ( the keyhole ) of the enzyme ( the lock ).

Compare: Induced fit model   See also: enzyme , active site , substrate

Lock-and-key vs. Induced Fit Model

At present, two models attempt to explain enzyme-substrate specificity; one of which is the lock-and-key model , and the other is the Induced fit model . The lock and key model theory was first postulated by  Emil Fischer   in 1894. The lock-and-key enzyme action proposes the high specificity of enzymes. However, it does not explain the stabilization of the transition state that the enzymes achieve. The induced fit model (proposed by Daniel Koshland in 1958) suggests that the active site continues to change until the substrate is completely bound to the active site of the enzyme, at which point the final shape and charge are determined. Unlike the lock-and-key model, the induced fit model shows that enzymes are rather flexible structures. Nevertheless, Fischer’s Lock and Key theory laid an important foundation for subsequent research, such as during the refinement of the enzyme-substrate complex mechanism, as ascribed in the induced fit model. The lock-and-key hypothesis has opened ideas where enzyme action is not merely catalytic but incorporates a rather complex process in how they interact with the correct substrates with precision.

Key Components

Components of the lock and key model:

  • Enzyme : the enzyme structure is a three-dimensional protein configuration, with an active site from where the substrate binds.
  • Substrate : often an organic molecule, a substrate possesses a structural feature that complements the geometry of the enzyme’s active site.

In the lock and key model, both the enzymes and the substrates facilitate the formation of a complex that lowers the activation energy needed for a chemical transformation to occur. Such reduction in the activation energy allows the chemical reaction to proceed at a relatively faster rate, making enzymes crucial in various biological and molecular processes.

Lock-and-key Model Examples

Some of the common examples that are often discussed in the context of the Lock and Key Model are as follows:

  • Enzyme lactate dehydrogenase with a specific active site for its substrates, pyruvate and lactate. The complex facilitates the interconversion of pyruvate and lactate during anaerobic respiration
  • Enzyme carbonic anhydrase with a specific active site for the substrates carbon dioxide and water. The complex facilitates the hydration of carbon dioxide, forming bicarbonate
  • Enzyme lysozyme binding with a bacterial cell wall peptidoglycan, which is a vital immune function

Choose the best answer. 

Send Your Results (Optional)

  • Aryal, S. and Karki, P. (2023).  “Lock and Key Model- Mode of Action of Enzymes”. Microbenotes.com. https://microbenotes.com/lock-and-key-model-mode-of-action-of-enzymes/
  • Farhana, A., & Lappin, S. L. (2023, May).  Biochemistry, Lactate Dehydrogenase . Nih.gov; StatPearls Publishing. https://www.ncbi.nlm.nih.gov/books/NBK557536/

©BiologyOnline.com. Content provided and moderated by Biology Online Editors.

Last updated on January 11th, 2024

You will also like...

Homeostasis of organism water regulation.

Osmoregulation is the regulation of water concentrations in the bloodstream, effectively controlling the amount of water..

Freshwater Communities & Plankton

Planktons are microscopic organisms that live suspended in aquatic habitats. There are two groups: the phytoplanktons an..

Origins of Life on Earth

Earth was created around 4.5 billion years ago and life began not long after. Primitive life likely possessed the elemen..

Biological Cell Defense

Organisms employ different strategies to boost its defenses against antigens. Humans have an immune system to combat pat..

Human Biology – Food and Digestion

This tutorial recognizes the importance of food as a source of energy that will fuel many biological processes. A good d..

Takahē (Porphyrio hochstetteri)

Meet the colorful takahē, an extremely rare flightless bird. Find out more about its unique features and why they matte..

Related Articles...

No related articles found

Microbe Notes

Microbe Notes

Lock and Key Model- Mode of Action of Enzymes

Enzymes are biological catalysts. These are commonly proteins but also include RNA (ribozymes) molecules that catalyze chemical reactions by lowering the activation energy of a reaction. These are known to speed up the rate of a reaction millions of times faster than the reaction without enzymes. Nearly all biological reactions require enzymes to transform substrate into products. The substrate is the reactant molecule upon which enzymes act during a chemical reaction, and products are the substances formed as a result of a chemical reaction. A single reactant molecule can decompose to give multiple products. Similarly, two reactants can enter into a reaction to yield products. These are reusable even after the completion of the reaction. Chemical properties such as charge and pH are vital in enzymatic reactions.

Binding between enzymes and reactant molecules takes place in such a way that chemical bond-breaking and bond-forming processes occur more readily. Meanwhile, no change in ∆the G value of a reaction takes place, thereby not altering the energy-releasing or energy-absorbing process of the reaction. However, it lowers the energy of the transition state, the topmost unstable state where the activated complex is formed from reactants that later give products.

Free Energy Changes (dG) in Exergonic and Endergonic Reactions

Table of Contents

Interesting Science Videos

Enzyme’s Active site and Substrate Specificity

Enzymes are relatively larger than the substrates, whose only a small fraction is involved in catalysis by reducing chemical activation energy, also known as the catalytic site, and the other portion for binding with the substrate and orienting them also known as the binding site. The catalytic site and binding site altogether form the active site of an enzyme. Usually, there are two active sites in an enzyme.

  •  The active site of enzymes is a cleft portion, composed of a small number of a unique combination of amino acid residues, usually three to four in number, which make up only ~10-20% of the volume of an enzyme. 
  • The remaining amino acids are used to maintain tertiary structure by proper scaffold folding through non-covalent interactions.
  • Non-covalent interaction between enzyme and substrate in correct orientation favors their reaction. These interactions include hydrogen bonds , hydrophobic bonds, ionic interactions, and Van der Waal’s interactions.
  • However, transient covalent bonds between enzymes and substrates are also formed during the time of reaction.
  • Side chains of amino acids play an important role in highly specific three-dimensional conformation at the level of the active site. These are large or small, hydrophilic or hydrophobic, acidic or basic.
  • The specific shape, size, and chemical behavior of enzymes are determined by the nature of amino acids and their 3D space in the active site.

Specificity is a distinctive feature of enzymes where they have a unique ability to choose an exact substrate from a group of similar chemical molecules. Their specificity towards their substrate varies to a different extent. These are of different types, namely: Bond specificity, Group specificity. Substrate specificity, Stereospecificity, Geometrical specificity, and Co-factor specificity.

Substrate specificity is also k/a absolute specificity for the enzyme’s specificity towards one substrate and one reaction. For e.g., Lactase acts on the B-1-4 glycosidic linkage of lactose to yield galactose and glucose. The restrictive nature of enzymes towards the choice of substrate can be attributed to the enzymatic activity of two oxidoreductase enzymes. Alcohol dehydrogenase uses its substrate alcohol while lactic acid dehydrogenase act on lactic acid. Although these two enzymes function with the mechanism of oxidation and reduction reaction, their substrates can’t be used interchangeably. This is because the different structure of each substrate prevents their fitting into the active site of the alternative enzyme.

In most cases, cofactors, the non-protein molecules, are required to ensure an efficient enzyme-facilitated chemical reaction. These function to bind with enzymes via either ionic interaction or covalent interactions. Metal ions (such as minerals) and co-enzymes (vitamin derivatives) are cofactors.

Lock and Key Model

A German scientist, Emil Fischer postulated the lock and key model in 1894 to explain the enzyme’s mode of action. Fischer’s theory hypothesized that enzymes exhibit a high degree of specificity towards the substrate. This model assumes that the active site of the enzyme and the substrate fit perfectly into one another such that each possesses specific predetermined complementary geometric shapes and sizes. Thus, the shape of the enzyme and substrate do not influence each other. This specificity is analog to the lock and key model, where the lock is the enzyme, and the key is the substrate. In certain circumstances, if a second substrate similar in shape and size to the primary substrate is made to bind to the enzyme, this second substrate also fits in the active site too.

How does Lock and Key Model work?

  • Binding of the substrate(s) to the enzyme at their active site takes place, thereby forming an enzyme-substrate complex.
  • Enzymes catalyze the chemical reaction to take place, which can either be a synthesis reaction (favors bond formation) or a decomposition reaction (favors bond breakage).
  • As a result, the formation of one or more products takes place, and the enzymes are released for their reuse in the next reaction.

Lock and Key Model

Limitations of Lock and Key Model

  • It doesn’t explain how the enzyme-substrate complex is stabilized in the transition state.
  • This model supposes the enzyme is a rigid structure whose shape does not change upon binding with a suitable substrate. However, this is not supported by recent research, which states that there is a change in conformation of the active site of the enzyme upon binding of substrate.
  • It does not describe the condition for binding multiple substrates to the enzyme.

Later, it was found that enzyme specificity toward one substrate is not always true. Although there are enzymes that specifically bind with only one substrate, there are also enzymes that exhibit broad specificity towards different but similarly structured substrates, such as lipase that can bind to different types of lipids. Similarly, proteases such as trypsin and chymotrypsin can degrade multiple types of proteins. Thus, the lock and key model is flawed, and the induced fit model was introduced to give a more refined view of the mode of enzymatic action.

  • Blanco, A., & Blanco, G. (2017). Medical Biochemistry. Academic Press. https://www.khanacademy.org/science/ap-biology/cellular-energetics/enzyme-structure-and-catalysis/a/enzymes-and-the-active-site
  • https://www.biologyonline.com/dictionary/substrate-specificity
  • https://www.britannica.com/science/protein/The-mechanism-of-enzymatic-action
  • https://www.biologyonline.com/dictionary/lock-and-key-model
  • https://study.com/learn/lesson/lock-key-model-vs-induced-fit-model.html
  • https://chem.libretexts.org/Bookshelves/Organic_Chemistry/Organic_Chemistry_(McMurry)/26%3A_Biomolecules-_Amino_Acids_Peptides_and_Proteins/26.11%3A_Enzymes_and_Coenzymes
  • https://en.wikibooks.org/wiki/Structural_Biochemistry/Protein_function/Lock_and_Key
  • https://ib.bioninja.com.au/higher-level/topic-8-metabolism-cell/untitled-6/models-of-action.html

About Author

Photo of author

Prakriti Karki

Leave a Comment Cancel reply

Save my name, email, and website in this browser for the next time I comment.

Algor Cards

Cosa ne pensi di noi?

Il tuo nome

La tua email

lock and key hypothesis of enzyme specificity

The Lock and Key Theory: Understanding Enzyme Specificity and Catalysis

Mappa concettuale.

lock and key hypothesis of enzyme specificity

The Lock and Key Theory, introduced by Emil Fischer, is a fundamental concept in biochemistry that explains enzyme specificity. It compares the enzyme's active site to a lock and the substrate to a key, illustrating how only the correct substrate can initiate a reaction. This theory is pivotal in understanding biochemical pathways, organic chemistry, and pharmaceuticals, influencing drug design by targeting enzyme active sites to treat diseases.

Mostra di più

Introduction to the Lock and Key Theory

Definition of the lock and key theory.

The Lock and Key Theory explains the specificity of enzyme action by comparing the enzyme's active site to a lock and the substrate to a key

Elements of the Lock and Key Theory

Enzyme and its active site

Enzymes are specialized proteins with a three-dimensional pocket, called the active site, designed to bind specific substrates

Substrates are molecules that bind to the enzyme's active site, triggering a specific reaction

Enzyme-substrate complex and product

The binding of the substrate to the active site forms an enzyme-substrate complex, which facilitates the conversion of the substrate into the product

Reaction sequence in the Lock and Key Theory

The Lock and Key Theory can be summarized by the reaction sequence \(E + S \rightarrow ES \rightarrow E + P\), where \(E\) represents the enzyme, \(S\) the substrate, \(ES\) the enzyme-substrate complex, and \(P\) the product

Applications of the Lock and Key Theory

Importance in biochemistry and pharmaceutical industry.

The Lock and Key Theory is a fundamental concept in biochemistry and is crucial in drug design for developing specific and effective pharmaceutical agents

Comparison with the Induced Fit Theory

The Induced Fit Theory complements the Lock and Key model by suggesting that the active site is dynamic and can adapt to accommodate slight variations in substrate structure

Educational resources for learning the Lock and Key Theory

Detailed diagrams and glossaries are valuable tools for understanding the step-by-step process of enzyme action and the key terms associated with the Lock and Key Theory

Significance of the Lock and Key Theory

Influence on our understanding of enzymatic function.

The Lock and Key Theory has greatly contributed to our understanding of enzymatic function and the complex chemical processes that sustain life

Practical applications in drug design

The Lock and Key Theory is essential in the development of enzyme inhibitors for treating diseases by disrupting pathological processes

Role in biochemistry education and research

The Lock and Key Theory remains a fundamental concept in biochemistry education and research, providing a starting point for exploring the complexities of enzyme behavior

Vuoi creare mappe dal tuo materiale?

Inserisci un testo, carica una foto o un audio su Algor. In pochi secondi Algorino lo trasformerà per te in mappa concettuale, riassunto e tanto altro!

Impara con le flashcards di Algor Education

Clicca sulla singola scheda per saperne di più sull'argomento.

lock and key hypothesis of enzyme specificity

Enzymes, which are specialized ______, work as catalysts and their action is explained by the ______ and Key Theory.

proteins Lock

lock and key hypothesis of enzyme specificity

Describe the active site of an enzyme.

Three-dimensional pocket on enzyme surface; binds substrate with high specificity; involves non-covalent interactions like hydrogen bonds, ionic interactions, van der Waals forces.

lock and key hypothesis of enzyme specificity

Explain the enzyme-substrate complex.

Transient complex formed when substrate binds to enzyme's active site; facilitates substrate's conversion into product.

lock and key hypothesis of enzyme specificity

Summarize the reaction sequence in the Lock and Key Theory.

E + S -> ES -> E + P; E is enzyme, S is substrate, ES is enzyme-substrate complex, P is product; enzyme remains unchanged post-reaction.

lock and key hypothesis of enzyme specificity

The ______ and Key Theory is crucial for understanding enzyme catalysis in organic chemistry and drug design in the pharmaceutical industry.

Originator of Induced Fit Theory

Daniel Koshland proposed the Induced Fit Theory in 1958.

Characteristic of enzyme's active site in Induced Fit Theory

Active site is dynamic, molds around substrate upon binding.

Role of substrate structure variability in enzyme specificity

Enzyme adaptability allows accommodation of substrates with slight structural variations.

Educational materials like ______ and glossaries are crucial for grasping the ______ Theory.

detailed diagrams Lock and Key

Lock and Key Theory - Basic Concept

Theory where enzymes and substrates fit together precisely like a lock and key, explaining enzyme specificity.

Enzyme Specificity - Importance

Critical for enzymes to catalyze only the correct reactions, ensuring proper metabolic function.

Drug Design - Lock and Key Relevance

Lock and Key Theory guides creation of enzyme inhibitors that mimic substrates, blocking unwanted reactions in disease.

Ecco un elenco delle domande più frequenti su questo argomento

Who introduced the lock and key theory and what does it explain about enzymes, what are the key components involved in the lock and key theory, how does the lock and key theory aid in the development of new pharmaceuticals, what is the difference between the lock and key and the induced fit theories of enzyme specificity, what educational tools are recommended for understanding the lock and key theory, despite advancements in biochemistry, why is the lock and key theory still significant, contenuti simili, esplora altre mappe su argomenti simili.

Accurate three-dimensional model of the human brain with visible left hemisphere, gloved hands supporting it on a neutral background.

Brain Development and Its Impact on Functioning

Three-dimensional molecular model of a protein with red alpha-helices, yellow beta sheets and colored side chains on a gradient background.

Amino Acids and Proteins

Molecular model of mRNA intertwined with tRNA molecules, with colored nitrogenous bases and amino acids attached, on a neutral background.

Messenger RNA and Protein Synthesis

High resolution 3D representation of hemoglobin molecule with quaternary structure, subunits in red, blue, green and purple, yellow heme groups.

Protein Structure and Function

Three-dimensional molecular model of a ribosome with major and minor subunits, colored spheres indicate atoms and bonds, mRNA strand runs through.

Ribosomal RNA and its Role in Protein Synthesis

Close-up of a glass Petri dish with bacterial colonies on agar in the laboratory, stainless steel instrument above, background with microscope and notebook.

Enzymes: Biological Catalysts for Life

Macro photography of a cell membrane with protruding glycolipids and branched carbohydrate chains in shades of blue and green on a blurred background.

Glycolipids: Structure, Function, and Importance

lock and key hypothesis of enzyme specificity

Non trovi quello che cercavi?

Cerca un argomento inserendo una frase o una parola chiave

The Fundamentals of Enzyme Specificity: Lock and Key Theory

Traditional metal key inserted halfway into a cylinder lock, with intricate design and brushed finish on neutral blurred background.

Exploring the Components and Dynamics of the Lock and Key Model

The role of the lock and key theory in organic chemistry and drug discovery, lock and key versus induced fit: diverse models of enzyme specificity, educational resources for grasping the lock and key theory, the enduring influence of the lock and key theory on biochemical science.

Modifica disponibile

Encyclopedia Britannica

  • History & Society
  • Science & Tech
  • Biographies
  • Animals & Nature
  • Geography & Travel
  • Arts & Culture
  • Games & Quizzes
  • On This Day
  • One Good Fact
  • New Articles
  • Lifestyles & Social Issues
  • Philosophy & Religion
  • Politics, Law & Government
  • World History
  • Health & Medicine
  • Browse Biographies
  • Birds, Reptiles & Other Vertebrates
  • Bugs, Mollusks & Other Invertebrates
  • Environment
  • Fossils & Geologic Time
  • Entertainment & Pop Culture
  • Sports & Recreation
  • Visual Arts
  • Demystified
  • Image Galleries
  • Infographics
  • Top Questions
  • Britannica Kids
  • Saving Earth
  • Space Next 50
  • Student Center

enzyme; active site

key-lock hypothesis

Learn about this topic in these articles:, chromatography.

elution chromatography

Very specific intermolecular interactions, “lock and key,” are known in biochemistry. Examples include enzyme-protein, antigen-antibody, and hormone-receptor binding. A structural feature of an enzyme will attach to a specific structural feature of a protein. Affinity chromatography exploits this feature by binding a

peptide

…and enzyme, called the “key–lock” hypothesis, was proposed by German chemist Emil Fischer in 1899 and explains one of the most important features of enzymes, their specificity. In most of the enzymes studied thus far, a cleft, or indentation, into which the substrate fits is found at the active…

The key–lock hypothesis ( see above The nature of enzyme-catalyzed reactions) does not fully account for enzymatic action; i.e., certain properties of enzymes cannot be accounted for by the simple relationship between enzyme and substrate proposed by the key–lock hypothesis. A theory called the induced-fit theory retains…

Fundamentals of Enzymatic Processes

  • First Online: 24 June 2017

Cite this chapter

lock and key hypothesis of enzyme specificity

  • R. Navanietha Krishnaraj 3 ,
  • Aditi David 3 &
  • Rajesh K. Sani 3  

732 Accesses

3 Citations

A basic and clear understanding about enzymes is essential for the readers before they begin to learn about the extremozymes (enzymes in extreme conditions). The chapter begins with the basic concepts of enzymes, roles of enzymes in biological systems, components of enzymes, detailed list of applications of the enzymes and the history of enzymology. The section on specificity of enzymes explains five different types of specificity namely Absolute Substrate specificity, Broad specificity (Group specificity), Bond specificity (Relative specificity), Stereochemical specificity and Reaction specificity. Specificity is the key characteristic of the enzyme and has crucial role in terms of selectivity and catalytic activity of the enzyme. The chapter covers the different classification methods of enzymes and enzyme nomenclature. The chapter gives a clear explanation about the mechanisms of enzyme with special emphasis on Lock and Key Theory and Induced Fit Hypothesis. Different units of enzyme activity (Katal, IU, Turnover number), different models of enzyme kinetics, types of enzymes inhibition and different strategies for immobilization of enzymes are addressed. Finally, the chapter describes the various applications of extremozymes.

This is a preview of subscription content, log in via an institution to check access.

Access this chapter

Subscribe and save.

  • Get 10 units per month
  • Download Article/Chapter or eBook
  • 1 Unit = 1 Article or 1 Chapter
  • Cancel anytime
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
  • Available as EPUB and PDF
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info
  • Durable hardcover edition

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

Similar content being viewed by others

lock and key hypothesis of enzyme specificity

Thermodynamics and Stability

lock and key hypothesis of enzyme specificity

Enzymes: Historical Aspects

Closure – whither enzymology.

Agrawal D, Patidar P, Banerjee T, Patil S (2004) Production of alkaline protease by Penicillium sp. under SSF conditions and its application to soy protein hydrolysis. Process Biochem 39:977–981

Article   CAS   Google Scholar  

Aurilia V, Parracino A, D'Auria S (2008) Microbial carbohydrate esterases in cold adapted environments. Gene 410:234–240

Article   CAS   PubMed   Google Scholar  

Buchholz K, Poulson PB (2000) Overview of history of applied biocatalysis. In: Straathof AJJ, Adlercreutz P (eds) Applied biocatalysis. Harwood Academic Publishers, Amsterdam

Google Scholar  

Bialkowska AM, Cieslinski H, Nowakowska KM, Kur J, Turkiewicz M (2009) A new beta-galactosidase with a low temperature optimum isolated from the Antarctic arthrobacter sp. 20B: gene cloning, purification and characterization. Arch Microbiol 191:825–835

Chiba S, Yamada M, Isobe K (2015) Novel acidophilic beta-galactosidase with high activity at extremely acidic pH region from Teratosphaeria acidotherma AIU BGA-1. J Biosci Bioeng 120:263–267

Collins T, Hoyoux A, Dutron A, Georis J, Genot B, Dauvrin T, Arnaut F, Gerday C, Feller G (2006) Use of glycoside hydrolase family 8 xylanases in baking. J Cereal Sci 43:79–84

Dahiya N, Tewari R, Hoondal GS (2006) Biotechnological aspects of chitinolytic enzymes: a review. Appl Microbiol Biotechnol 71:773–782

Devi SG, Fathima AA, Sanitha M, Iyappan S, Curtis WR, Ramya M (2016) Expression and characterization of alkaline protease from the metagenomic library of tannery activated sludge. J Biosci Bioeng 122:694–700

Fujinami S, Fujisawa M (2010) Industrial applications of alkaliphiles and their enzymes – past, present and future. Environ Technol 31:845–856

Gupta R, Beg QK, Lorenz P (2002) Bacterial alkaline proteases: molecular approaches and industrial applications. Appl Microbiol Biotechnol 59:15–32

Guy JE, Isupov MN, Littlechild JA (2003) The structure of an alcohol dehydrogenase from the hyperthermophilic Archaeon aeropyrum pernix. J Mol Biol 331:1041–1051

Haddar A, Agrebi R, Bougatef A, Hmidet N, Sellami-Kamoun A, Nasri M (2009) Two detergent stable alkaline serine-proteases from Bacillus mojavensis A21: purification, characterization and potential application as a laundry detergent additive. Bioresour Technol 100:3366–3373

Hildebrandt P, Wanarska M, Kur J (2009) A new cold-adapted beta-D-galactosidase from the Antarctic arthrobacter sp. 32c – gene cloning, overexpression, purification and properties. BMC Microbiol 9:151

Article   PubMed   PubMed Central   Google Scholar  

Horikoshi K (1999) Alkaliphiles: some applications of their products for biotechnology. Microbiol Mol Biol Rev 63:735–750

CAS   PubMed   PubMed Central   Google Scholar  

Joseph B, Ramteke PW, Thomas G (2008) Cold active microbial lipases: some hot issues and recent developments. Biotechnol Adv 26:457–470

Liao H, Xu C, Tan S, Wei Z, Ling N, Yu G, Raza W, Zhang R, Shen Q, Xu Y (2012) Production and characterization of acidophilic xylanolytic enzymes from Penicillium oxalicum GZ-2. Bioresour Technol 123:117–124

Payen A, Persoz JF (1833) Mémoire su la diastase, les principaux produits de ses réactions, et leurs applications aux arts industriels. Ann Chim Phys 53:73–92

Phadatare SU, Deshpande VV, Srinivasan MC (1993) High activity alkaline protease from Conidiobolus coronatus (NCL 86.8.20): enzyme production and compatibility with commercial detergents. Enzyme Microb Technol 15:72–76

Sinha R, Srivastava AK, Khare SK (2014) Efficient proteolysis and application of an alkaline protease from halophilic Bacillus sp. EMB9. Prep Biochem Biotechnol 44:680–696

Sumner JB, Somers GF (1953) Chemistry and methods of enzymes, third edition. Soil Sci 76(2):166

Article   Google Scholar  

Toogood HS, Hollingsworth EJ, Brown RC, Taylor IN, Taylor SJ, McCague R, Littlechild JA (2002) A thermostable L-aminoacylase from Thermococcus litoralis: cloning, overexpression, characterization, and applications in biotransformations. Extremophiles 6:111–122

Toogood HS, Brown RC, Line K, Keene PA, Taylor SJC, McCague R, Littlechild JA (2004) The use of a thermostable signature amidase in the resolution of the bicyclic synthon (rac)-γ-lactam. Tetrahedron 60:711–716

Tutino ML, di Prisco G, Marino G, de Pascale D (2009) Cold-adapted esterases and lipases: from fundamentals to application. Protein Pept Lett 16:1172–1180

Ueda M, Goto T, Nakazawa M, Miyatake K, Sakaguchi M, Inouye K (2010) A novel cold-adapted cellulase complex from Eisenia foetida: characterization of a multienzyme complex with carboxymethylcellulase, beta-glucosidase, beta-1,3 glucanase, and beta-xylosidase. Comp Biochem Physiol B Biochem Mol Biol 157:26–32

Article   PubMed   Google Scholar  

Wang F, Hao J, Yang C, Sun M (2010) Cloning, expression, and identification of a novel extracellular cold-adapted alkaline protease gene of the marine bacterium strain YS-80-122. Appl Biochem Biotechnol 162:1497–1505

Wang C, Zhang J, Wang Y, Niu C, Ma R, Wang Y, Bai Y, Luo H, Yao B (2016) Biochemical characterization of an acidophilic beta-mannanase from Gloeophyllum trabeum CBS900.73 with significant transglycosylation activity and feed digesting ability. Food Chem 197:474–481

Download references

Author information

Authors and affiliations.

Department of Chemical and Biological Engineering, South Dakota School of Mines and Technology, 501 East St. Joseph Street, Rapid City, SD, 57701-3995, USA

R. Navanietha Krishnaraj, Aditi David & Rajesh K. Sani

You can also search for this author in PubMed   Google Scholar

Corresponding author

Correspondence to Rajesh K. Sani .

Editor information

Editors and affiliations.

Department of Chemical and Biological Engineering, South Dakota School of Mines and Technology, Rapid City, South Dakota, USA

Rajesh K. Sani

R. Navanietha Krishnaraj

Rights and permissions

Reprints and permissions

Copyright information

© 2017 Springer International Publishing AG

About this chapter

Navanietha Krishnaraj, R., David, A., Sani, R.K. (2017). Fundamentals of Enzymatic Processes. In: Sani, R., Krishnaraj, R. (eds) Extremophilic Enzymatic Processing of Lignocellulosic Feedstocks to Bioenergy. Springer, Cham. https://doi.org/10.1007/978-3-319-54684-1_2

Download citation

DOI : https://doi.org/10.1007/978-3-319-54684-1_2

Published : 24 June 2017

Publisher Name : Springer, Cham

Print ISBN : 978-3-319-54683-4

Online ISBN : 978-3-319-54684-1

eBook Packages : Biomedical and Life Sciences Biomedical and Life Sciences (R0)

Share this chapter

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

  • Publish with us

Policies and ethics

  • Find a journal
  • Track your research

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
  • My Bibliography
  • Collections
  • Citation manager

Save citation to file

Email citation, add to collections.

  • Create a new collection
  • Add to an existing collection

Add to My Bibliography

Your saved search, create a file for external citation management software, your rss feed.

  • Search in PubMed
  • Search in NLM Catalog
  • Add to Search

How Emil Fischer was led to the lock and key concept for enzyme specificity

Affiliation.

  • 1 Department of Chemistry, University of Alberta, Edmonton, Canada.
  • PMID: 7942253

PubMed Disclaimer

Similar articles

  • Emil Fischer--unequalled classicist, master of organic chemistry research, and inspired trailblazer of biological chemistry. Kunz H. Kunz H. Angew Chem Int Ed Engl. 2002 Dec 2;41(23):4439-51. doi: 10.1002/1521-3773(20021202)41:23 3.0.CO;2-6. Angew Chem Int Ed Engl. 2002. PMID: 12458504 No abstract available.
  • Biocatalysis in organic chemistry (Part I): Past and present. Turner MK. Turner MK. Trends Biotechnol. 1995 May;13(5):173-7. doi: 10.1016/S0167-7799(00)88934-9. Trends Biotechnol. 1995. PMID: 7540395
  • Emil Fischer (1852-1919)--organic chemist. [No authors listed] [No authors listed] JAMA. 1966 Oct 17;198(3):313-4. JAMA. 1966. PMID: 5330444 No abstract available.
  • Beginnings of microbiology and biochemistry: the contribution of yeast research. Barnett JA. Barnett JA. Microbiology (Reading). 2003 Mar;149(Pt 3):557-567. doi: 10.1099/mic.0.26089-0. Microbiology (Reading). 2003. PMID: 12634325 Review.
  • A history of research on yeasts 7: enzymic adaptation and regulation. Barnett JA. Barnett JA. Yeast. 2004 Jul 15;21(9):703-46. doi: 10.1002/yea.1113. Yeast. 2004. PMID: 15282797 Review. No abstract available.
  • The Conformational Contribution to Molecular Complexity and Its Implications for Information Processing in Living Beings and Chemical Artificial Intelligence. Gentili PL. Gentili PL. Biomimetics (Basel). 2024 Feb 19;9(2):121. doi: 10.3390/biomimetics9020121. Biomimetics (Basel). 2024. PMID: 38392167 Free PMC article.
  • Editorial: Intrinsically Disordered Proteins and Regions: The Challenge to the Structure-Function Relationship. Toto A, Sormanni P, Paissoni C, Uversky VN. Toto A, et al. Front Mol Biosci. 2022 Jul 6;9:962643. doi: 10.3389/fmolb.2022.962643. eCollection 2022. Front Mol Biosci. 2022. PMID: 35874612 Free PMC article. No abstract available.
  • The synapse: people, words and connections. Tansey EM. Tansey EM. Neuronal Signal. 2022 Jun 8;6(2):NS20220017. doi: 10.1042/NS20220017. eCollection 2022 Jun. Neuronal Signal. 2022. PMID: 35813266 Free PMC article.
  • A Comprehensive Insight into Fungal Enzymes: Structure, Classification, and Their Role in Mankind's Challenges. El-Gendi H, Saleh AK, Badierah R, Redwan EM, El-Maradny YA, El-Fakharany EM. El-Gendi H, et al. J Fungi (Basel). 2021 Dec 28;8(1):23. doi: 10.3390/jof8010023. J Fungi (Basel). 2021. PMID: 35049963 Free PMC article. Review.
  • The future of biomolecular simulation in the pharmaceutical industry: what we can learn from aerodynamics modelling and weather prediction. Part 1. understanding the physical and computational complexity of in silico drug design. Edwards T, Foloppe N, Harris SA, Wells G. Edwards T, et al. Acta Crystallogr D Struct Biol. 2021 Nov 1;77(Pt 11):1348-1356. doi: 10.1107/S2059798321009712. Epub 2021 Oct 27. Acta Crystallogr D Struct Biol. 2021. PMID: 34726163 Free PMC article.

Publication types

  • Search in MeSH

Personal name as subject

Linkout - more resources, full text sources.

  • Elsevier Science
  • Citation Manager

NCBI Literature Resources

MeSH PMC Bookshelf Disclaimer

The PubMed wordmark and PubMed logo are registered trademarks of the U.S. Department of Health and Human Services (HHS). Unauthorized use of these marks is strictly prohibited.

Captcha Page

We apologize for the inconvenience...

To ensure we keep this website safe, please can you confirm you are a human by ticking the box below.

If you are unable to complete the above request please contact us using the below link, providing a screenshot of your experience.

https://ioppublishing.org/contacts/

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

The PMC website is updating on October 15, 2024. Learn More or Try it out now .

  • Advanced Search
  • Journal List
  • Biol Direct

Logo of bioldirect

Conserved sequence motifs in human TMTC1, TMTC2, TMTC3, and TMTC4, new O-mannosyltransferases from the GT-C/PMT clan, are rationalized as ligand binding sites

Birgit eisenhaber.

1 Bioinformatics Institute (BII), Agency for Science, Technology and Research (A*STAR), 30 Biopolis Street, #07-01 Matrix, Singapore, 138671 Republic of Singapore

2 Genome Institute of Singapore (BII), Agency for Science, Technology and Research (A*STAR), 60 Biopolis Street, Singapore, 138672 Republic of Singapore

Swati Sinha

Chaitanya k. jadalanki, vladimir a. shitov.

3 Siberian State Medical University, Moskovskiy Trakt, 2, Tomsk, Tomsk Oblast 634050 Russia

Qiao Wen Tan

4 School of Biological Science (SBS), Nanyang Technological University (NTU), 60 Nanyang Drive, Singapore, 637551 Republic of Singapore

Fernanda L. Sirota

Frank eisenhaber, associated data.

All supporting data are submitted in Supplementary materials (Additional Files 1 , 2 , 3 , 4 and 5 ).

The human proteins TMTC1, TMTC2, TMTC3 and TMTC4 have been experimentally shown to be components of a new O-mannosylation pathway. Their own mannosyl-transferase activity has been suspected but their actual enzymatic potential has not been demonstrated yet. So far, sequence analysis of TMTCs has been compromised by evolutionary sequence divergence within their membrane-embedded N-terminal region, sequence inaccuracies in the protein databases and the difficulty to interpret the large functional variety of known homologous proteins (mostly sugar transferases and some with known 3D structure).

Evolutionary conserved molecular function among TMTCs is only possible with conserved membrane topology within their membrane-embedded N-terminal regions leading to the placement of homologous long intermittent loops at the same membrane side. Using this criterion, we demonstrate that all TMTCs have 11 transmembrane regions. The sequence segment homologous to Pfam model DUF1736 is actually just a loop between TM7 and TM8 that is located in the ER lumen and that contains a small hydrophobic, but not membrane-embedded helix. Not only do the membrane-embedded N-terminal regions of TMTCs share a common fold and 3D structural similarity with subgroups of GT-C sugar transferases. The conservation of residues critical for catalysis, for binding of a divalent metal ion and of the phosphate group of a lipid-linked sugar moiety throughout enzymatically and structurally well-studied GT-Cs and sequences of TMTCs indicates that TMTCs are actually sugar-transferring enzymes. We present credible 3D structural models of all four TMTCs (derived from their closest known homologues 5ezm/5f15) and find observed conserved sequence motifs rationalized as binding sites for a metal ion and for a dolichyl-phosphate-mannose moiety.

Conclusions

With the results from both careful sequence analysis and structural modelling, we can conclusively say that the TMTCs are enzymatically active sugar transferases belonging to the GT-C/PMT superfamily. The DUF1736 segment, the loop between TM7 and TM8, is critical for catalysis and lipid-linked sugar moiety binding. Together with the available indirect experimental data, we conclude that the TMTCs are not only part of an O-mannosylation pathway in the endoplasmic reticulum of upper eukaryotes but, actually, they are the sought mannosyl-transferases.

Supplementary Information

The online version contains supplementary material available at 10.1186/s13062-021-00291-w.

The family of TMTC (transmembrane (TM) and tetratricopeptide (TPR) repeat-containing) proteins in human is represented by four paralogues: TMTC1 (isoform X3 with accession XP_016875493, 875 residues (AA); see comment below why sequence Q8IUR5 (882 AA) appears doubtful), TMTC2 (Q8N394, 830 AA), TMTC3 (Q6ZXV5, 915 AA) and TMTC4 (Q5T4D3, 741 AA). Their common sequence architecture consists of an N-terminal segment with transmembrane regions and intermittent loops and a C-terminal stretch of multiple, in the order of 10 TPR repeats.

After having long been genes with unknown function, first functional information trickled in from genome-wide association (GWAS) and family (FS) studies that linked TMTCs with neurological/psychiatric diseases, sensory organ disorders but also with other conditions. Although an original, GWAS-based claim for TMTC2 in primary open-angle glaucoma in a Japanese cohort [ 1 ] could not be confirmed in several follow-up studies (for Afro-Caribbean [ 2 ], Chinese [ 3 ], Japanese [ 4 ], Korean [ 5 ], Saudi Arabian [ 6 ] and South Indian [ 7 ] cohorts), new GWAS evidence for another set of single nucleotide polymorphisms in cohorts of mixed ethnic origin reemphasizes the link [ 8 ]. Optic disc area is influenced by TMTC2 in cohorts of European and Asian ancestry [ 9 ].

TMTC1 has been related to schizophrenia (via GWAS [ 10 ]) and it is differentially expressed in inflammatory bowel disease linked arthritis [ 11 ]. The circular RNA circTMTC1 inhibits skeletal muscle satellite cell differentiation in chicken [ 12 ]. TMTC2 is associated with non-syndromic sensorineural hearing loss (SNHL; via both GWAS and FS [ 13 , 14 ]). TMTC2 interactions with certain miRNAs hint towards a role in Parkinson’s disease [ 15 ]. GWAS associates TMTC2 with obesity in Caribean Hispanics [ 16 ] and Han Chinese [ 17 ], left ventricular mass increase [ 16 ] as well as with immune conditions such as eczema, asthma and ‘atopic march’ [ 18 ]. Family studies show TMTC3 mutations being causative for cobblestone lissencephaly [ 19 ] and periventricular nodular heterotopia with intellectual disability and epilepsy [ 20 ]. Genetic inactivation of TMTC4 in mice causes rapid, early postnatal cochlear hair cell death, leading to hearing loss [ 21 ]. TMTC4’s role in influencing bone mineral density is known from a transcriptome-wide association study [ 22 ].

Hence, the diversity of clinical effects hints towards human TMTCs having, most likely, very basic molecular and cellular functions with pleiotropic, context-specific effects. TMTC1 [ 23 ], TMTC2 [ 23 ], TMTC3 [ 24 , 25 ] and TMTC4 [ 21 , 25 ] were found to be located in the endoplasmic reticulum (ER). For all TMTCs, the TPR-containing C-terminal segment was shown to be located in the ER lumen (TMTC1/2 [ 23 ], TMTC3/4 [ 25 ]). TMTC1/2 were associated with intracellular calcium homeostasis [ 21 , 23 ]. TMTC3 was reported to have a potential role in ER stress response [ 24 ], TMTC4 was linked with unfolded protein response [ 21 ].

Dramatic progress in understanding TMTC function was recently achieved by Danish researchers collaborating with several American groups [ 25 , 26 ]. Knockout of all four TMTCs in HEK293 cells abolished O-mannosylation of a variety of cadherin and proto-cadherin proteins; thus, the TMTCs are members of a new O-mannosylation pathway that selectively processes cadherin-like targets [ 26 ]. Apparently, the presence of various TMTCs affects the spectrum of modified cadherins since the selective TMTC1/3 knockout (with TMTC2/4 remaining functional) produces a larger set of O-mannosyl glycopeptides in the mass-spectrometric analysis [ 26 ]. Further, TMTC3 complementation at the background of a combined four TMTC knockout in HEK293 cells rescues the O-mannosylation of E-cadherin and enhances cellular adherence [ 25 ]. TMTC3/4 knockdowns were demonstrated to delay gastrulation in frog [ 25 ]. Three known TMTC3 disease mutations in the N-terminal protein half (H67D, R71H, G384E) were shown to exhibit reduced protein half-life despite native ER localization.

Having followed the TMTC story since 2012, we were puzzled by the difficulties to consistently interpret the sequence-analytic findings in terms of biological function, a problem so nicely summarized by Larsen, Graham et al. [ 25 – 28 ]. It starts with something apparently simple such as the largely varying predicted transmembrane region (TM) numbers for various TMTCs due to evolutionary sequence divergence within their membrane-embedded N-terminal region and it does not end with the diversity of enzymatic activities and substrates of homologous proteins (largely sugar transferases), sometimes even with known 3D structure. In this work, we explore:

  • (i) To which extent can the sequence architecture of TMTCs be unified, especially with regard to their number of TMs?
  • (ii) What is the nature of the sequence segment homologous to Pfam model DUF1736?
  • (iii) Can the conservation of sequence motifs among TMTCs and known homologous sugar transferases (including those with known 3D structure) be rationalized in terms of catalysis and ligand/substrate binding?

If not otherwise mentioned, all sequence-analytic operations were carried out with the ANNOTATOR software suite [ 29 , 30 ], an in-house tool developed over ca. 20 years that integrates more than 40 academic tools (either self-programmed or used with permission of the original authors) for the prediction of protein structural and functional features. In the context of this work, the battery of programs for prediction of transmembrane regions, cellular export signals and for sequence similarity searches were especially important. In cases where completeness and recent updates of sequence and domain databases were critical, selected locally executed similarity searches were repeated on the respective websites supported by the original authors (BLAST [ 31 , 32 ], HHpred [ 33 , 34 ]) to make sure that no important hit from recent database additions was omitted.

Structural modelling of TMTCs by homology was carried out with Modeller (version 9.4) [ 35 ]. As it became clear during the subsequent analyses that the TMTCs harbor a binding site for a lipid-linked sugar, we used the Schrodinger suite [ 36 ] for the placement of this ligand. Subsequent induced fit relaxation and energy optimization of the complex followed published procedures [ 36 – 42 ].

Collection and sequence architecture of the TMTC1/2/3/4 superfamily

Pairwise similarity searches using the BLAST tool [ 31 , 32 ] and starting with any of the full-length human TMTC1, TMTC2, TMTC3 and TMTC4 sequences conveniently gather the superfamily of true TMTC orthologues in upper Eukarya and of TMTC-like proteins in other organisms including many hypothetical proteins, if at all, automatically annotated by sequence similarity.

The sequence architecture of human TMTCs is two-partite with an N-terminal segment consisting of transmembrane regions and intermitting loops (456 AA for N-TMTC1, 475 AA for N-TMTC2, 426 AA for N-TMCT3 and 462 AA for N-TMTC4) and a remaining C-terminal part comprising TPR repeats. This result was obtained by analysing human TMTC1/2/3/4 within the ANNOTATOR environment [ 29 , 30 ]. We applied the suite of transmembrane prediction tools (DAS-tmfilter [ 43 , 44 ], HMMTOP [ 45 , 46 ], PHOBIUS [ 47 , 48 ], TMHMM [ 49 , 50 ] and TOPPRED2 [ 51 , 52 ]) as well as comparisons with protein domain and protein repeat databases (PFAM [ 53 ], SMART [ 54 ], Miguel Andrade’s repeats [ 55 ]) via HMM searches [ 56 , 57 ].

When we repeat the simple BLAST searches with just these N-terminal segments of TMTC1/2/3/4, apparently the same superfamily of TMTCs is collected (in the order of ~ 10,000 hits with E-value < 3.e-4 and above 60% query sequence coverage; details not shown). Phylogenetically, true TMTC orthologues and TMTC-like proteins are found throughout the eukaryote kingdom with homologues even among prokaryotes but the set of four paralogues per organism with full coverage of the N-terminal domain can be systematically detected only from vertebrates down to the insect level. Already in the complete genome of the worm Caenorhabditis elegans , just two TMTCs are known (TMTC1: Q20144/NP_509123, TMTC2: NP_504200).

We created a grand alignment of the full set of the N-terminal segments of TMTCs from six animal organisms ( Homo sapiens , Bos taurus , Gallus gallus , Xenopus laevis , Danio rerio , Drosophila melanogaster ; see Fig.  1 and Additional file  1 ) to study family-specific and superfamily-wide sequence conservation patterns.

An external file that holds a picture, illustration, etc.
Object name is 13062_2021_291_Fig1_HTML.jpg

Grand alignment of N-terminal segments of TMTCs together with sequences of selected sugar transferases with known 3D structure. We show the grand alignment of the full set of the N-terminal segments of TMTCs from six organisms ( Homo sapiens (Hs), Bos taurus (Bt), Gallus gallus (Gg), Xenopus laevis (Xl), Danio rerio (Dr), Drosophila melanogaster (Dm)) together with the sequences taken from 5ezm chain A [ 58 ], 6s7t chain A [ 59 ], 5ogl chain A [ 60 ] and 6p25 chain A [ 59 ]. For supporting navigation in the alignment, the location of the TMs in human TMTC1 and in 5ezm are shown. The figure was generated with Jalview [ 61 ] using an externally created and manually edited multiple alignment (in the SEAVIEW environment [ 62 , 63 ]). The location of TMs in TMTC1 follows the observations from the 3D structural model created in the course of this work and, at some places, does differ slightly from the sequence-analytic predictions provided in Additional File 2 . The following sequence segments have been excluded from the alignment and replaced by “XX”: in TMTC1_B, 244–304 after TM6; in TMTC1_Gg, 251–310 after TM6; in TMTC1_Dm, 358-417after TM8; in TMTC2_Hs, 337–393 after TM8; in TMTC2_Bt (G3MY32_BOVIN), 334–393 after TM8; in TMTC2_Gg (F1NPM4_CHICK), 324–380 after TM8; in TMTC2_Xl, 337–393 after TM8; in TMTC2_Dr (F1R0Y9_DANRE), 346–401 after TM8, in TMTC2_Dm, 360–504 after TM8; in 6S7T, 288–348 after TM6 and 486–535 after TM10; in 6P25, 219–261 after TM6, 312–531 after TM7 and 560–585. Please note that, as result of the excluded sequence stretches in some sequences, the residue numbering in the figure might deviate from the residue numbering in the respective entry of the sequence database. Additional information for this figure is provided in Additional Files 1 and 2 available with this article. For locating specific residues in the alignment, we recommend first finding the nearby TMs and then looking for conserved motifs next to them

As a first goal during the alignment creation, we wanted to understand the number and sequence localization of TM regions in the human TMTCs. In the literature, the number of TM regions in the N-terminal segment of various human TMTCs is reported to be different for various TMTCs and between 8 and 12 [ 25 – 28 ]. The confusion is not surprising as TM region predictors behave erratically in the twilight range of their scoring function [ 43 ]. Just one additional polar residue can bring the hydrophobicity of the candidate sequence segment below the threshold. And the boundaries of TM regions are typically heuristically determined bringing the length near 20 residues.

This variation of TM region number among TMTCs is potentially conflicting with evolutionarily conserved function as the latter requires homologous loop segments being located in the same subcellular space (in the ER or in the cytoplasm). Thus, membrane topology needs to be conserved among species within a given TMTC family and, to a large extent, also among various TMTC paralogues. As a further constraint, the C-terminal, TPR-comprising region is shown to be located in the ER for all TMTCs [ 23 , 25 ].

For all 24 sequences in Fig. ​ Fig.1, 1 , locations of potential TM regions were identified with the full suite of the five TM predictors in the ANNOTATOR [ 29 , 30 ]. In total, we find 12 regions with hydrophobic motifs that are predicted as TM regions in at least some sequences for three out of four families TMTC1, TMTC2, TMTC3 and TMTC4 (see Additional File  2 ). Four major discrepancies and issues are observed:

  • The most N-terminal TM region might actually be a signal peptide.
  • In the human TMTC1 sequence as in Q8IUR5, there is no hit for TM7. But it does exist in the sequence version of TMTC1 with accession XP_016875493 (isoform X3).
  • In human TMTC3, TM3 is only weakly recognized.
  • All TMTC sequences have a segment with significant sequence similarity to the Pfam domain DUF1736 (E-value < 1.e-30 for any of the human TMTCs in an HMMER search against Pfam-A [ 53 ]). The TM segment predictors suggest a TM region inside this segment for all human TMTCs except for TMTC2.

First, the most N-terminal hydrophobic region in all human TMTCs seems to be a true TM segment, maybe, a signal anchor but not a signal peptide as the sequence assessments with SIGNALP version 5 [ 64 ] show. The following loop contains the strongly conserved DD motif that, if having an enzymatic function, needs to be localized in the ER. Consequently, the N-terminus of TMTCs appears cytoplasmic. With the C-terminus in the ER, TMTCs need to have an uneven number of TM regions so that the TPR segment can reside inside the ER lumen [ 23 , 25 ].

Second, we encountered serious difficulties when attempting to include the canonical TMTC1 sequence Q8IUR5 into the grand alignment, especially in the region that includes TM7 and the DUF1736 hit (which is much worse in Q8IUR5 with E-value=3.e-19 compared with other TMTCs). This would not have surprised anyone if the sequence were from a more obscure insect or fish genome but Q8IUR5 is a human protein. Searching human sequences with TMTC1 from Bos taurus or Gallus gallus delivers XP_016875493 (TMTC1 isoform X3) as the sequence that can be much easier aligned with TMTC1s from other species as well as with other TMTCs. At the same time, searching the Bos taurus or Gallus gallus proteomes with human Q8IUR5 does not deliver a better, more similar isoform than the best homologue found with XP_016875493. Thus, it cannot be excluded that Q8IUR5 has sequence errors in the region 245–312 (with the corresponding region 245–305 in XP_016875493 being the correct version). While none of the five TM region predictors finds a trace of a hit for TM7 in Q8IUR5, it is confidently predicted by the majority of them in XP_016875493.

Third, the evolutionary argument (see Fig. ​ Fig.1) 1 ) strongly suggests that the respective regions for TM3 in human TMTC3 are just subthreshold for the TM predictors (compared with other human TMTCs, there are additional polar residues (Ser119, Ser120 and Ser124) in the respective sequence KSSVIASLLFAVHPIHT (residues 118–134) of human TMTC3).

Fourth, the sequence segment predicted to be a TM region as part of the DUF1736 hit is actually not membrane-embedded. When checking the TMTCs against sequences with known 3D structures via HHpred [ 33 , 34 ] as implemented in the ANNOTATOR environment [ 29 , 30 ], we find convincing statistically significant similarity of the N-terminal portions of TMTCs to structures such as 5ezm [ 58 ]. For example in the case of N-TMTC1, the E-value is 1.9e-22. Comparison with the alignment delivered by HHpred reveals that the segment FPNFFFI (261–267 in 5ezm), a small, quite hydrophobic helix at the ER side and with its axis parallel to the membrane, aligns with the segment 318–324 in human TMTC1. Notably, the segment 311–324 is the common core from TM predictions by four different TM predictors (TMHMM, PHOBIUS, DAS-tmfilter, and HMMTOP). Similar observations are available in other homologous structures. TMTC1’s segment 311–324 hits the same type of small, hydrophobic helix in the ER lumen parallel to the membrane in 5ogl (found with E-value 2.7e-15 by HHpred; segment 325–333 with sequence PEVFMQRIS [ 60 ]) or in 6s7t (found with E-value 2.4e-17 by HHpred; segment 382–389 with sequence GRFYSLWD [ 65 ]).

Thus, we can convincingly conclude (i) that the DUF1736-similar region in TMTCs, actually just a loop between TM7 and TM8 located in the endoplasmic reticulum lumen, does not contain a TM region, (ii) that all human TMTCs comprise 11 TM regions in their N-terminal sequence portion and (iii) that the N-terminus is located in the cytoplasm and the C-terminal TPR domain is in the ER lumen (see also Fig.  2 ).

An external file that holds a picture, illustration, etc.
Object name is 13062_2021_291_Fig2_HTML.jpg

Cartoon of the membrane topology of the N-terminal domain of TMTCs and localization of important substructures and residues. The figure shows schematic representation of the overall structural elements and the connectivity of TMTCs. The TM helices are shown in yellow cylinders and marked as I to XI while the helical regions in the lumen are shown in green cylinders and are marked as JM1, JM2 and JM3. The lumenal loops are numbered from EL1 to EL5. The whole TPR region is shown as a single block colored in cyan. The figure also highlights important residues which are (i) the strictly conserved DD motif (M1, Table ​ Table4) 4 ) in EL1 (loop between TM1 and TM2), (ii) conserved SHKSYRP motif (M2, Table ​ Table4) 4 ) also present in EL1, (iii) conserved lysine residue of KET(Q) xxT motif (M4, Table ​ Table4) 4 ) that forms a salt bridge with the phosphate group of DPM, (iv) glutamate residue from conserved KET(Q) xxT motif (M4, Table ​ Table4) 4 ) in EL3 and aspartate residue of the conserved DW motif (M4, Table ​ Table4) 4 ) in EL4, (v) strictly conserved arginine residue from conserved ERxxY motif (M7, Table ​ Table4) 4 ) in loop EL5 between TM9 and T10. All the important residues are colored in yellow except the metal binding residues which are highlighted in pink. The sequence position numbering corresponds to TMTC1. The location of TMs in TMTC1 follows the observations from the 3D structural model created in the course of this work and, at some places, does differ slightly from the sequence-analytic predictions provided in Additional File 2

Further, we wish to emphasize that the TM regions in TMTCs are largely of the complex type (the only consistently simple TMs are TM7 in TMTC3 from various species (data not shown)) [ 66 , 67 ]; thus, their sequences contain evolutionary information beyond the generally not informative hydrophobic background (sprinkled-in polar residues, glycine and proline are typically rare in TMs [ 68 , 69 ]) useful for sequence comparison in homology searches [ 70 – 72 ].

As mentioned by a reviewer, membrane topology prediction for proteins with TM regions has been attempted directly from sequence, typically following the TM segment prediction part [ 45 , 46 , 73 ]. As a trend, these prediction tools support the topology conclusions for the TMTCs but not always. For example, the probability for the N-terminus to be cytoplasmic was predicted by TMHMM [ 49 , 50 ] as follows: TMTC1 0.61, TMTC2 0.64, TMTC3 0.89, TMTC4 0.30. We think that the predicted number of TM regions (especially their even/uneven number) critically influences the correctness of the topology prediction. For TMTC1/2/3, nine TM regions were found by TMHMM (uneven as in the case of the actual 11 TM regions) but this number was predicted ten for TMTC4.

TMTCs are homologous to membrane-bound sugar transferases with known 3D structures

We summarized the findings related to the top hits of the HHpred searches with the N-TMTC1, N-TMTC2, N-TMTC3 and N-TMTC4 sequence segments in Table  1 . The original HHpred outputs are available as supplementary material (Additional File  3 ). All the hits have excellent E-values (<< 1.e-10) despite low sequence identities of the respective sequence alignments (all values between 8 and 13%; e.g., TMTC1/2/3/4 align with 5ezm with sequence identities 8, 13, 10 and 12% in the HHpred-generated alignments respectively); thus, the match of the physico-chemical property pattern between the respective sequences is excellent, especially for the TM segments and some loop regions next to them.

HHPred search with the N-terminal part of the four human TMTCs against PDB (PDB_mmCIF70_29_May, version 29/05/2020)

PDB IDN-TMTC1
(1–456)
N-TMTC2
(1–475)
N-TMTC3
(1–426)
N-TMTC4
(1–462)

5ezmA/5f15A

(578 AA)

[ ]

1.9E-225.9E-191.3E-214.2E-21
Q: 1–456Q: 1–475Q: 2–424Q: 1–460
T: 7–399T: 33–400T: 27–395T: 12–395

6s7tA

(826 AA)

[ ]

1.8E-172.3E-153.1E-171.2E-16
Q: 1–456Q: 2–475Q: 2–425Q: 1–462
T: 48–560T: 75–559T: 70–558T: 53–559

6s7oA

(705 AA)

[ ]

2.2E-174.9E-153.4E-174.1E-17
Q: 8–456Q: 2–475Q: 1–426Q: 1–461
T: 1–479T: 21–480T: 12–477T: 1–476

6eznF

(718 AA)

[ ]

1.4E-173E-155.1E-171.7E-16
Q: 9–454Q: 1–474Q: 2–424Q: 3–461
T: 1–467T: 19–469T: 14–465T: 1–466

3wajA

(875 AA)

[ ]

2.6E-178.6E-151.8E-171E-15
Q: 12–456Q: 1–474Q: 1–425Q: 7–459
T: 1–490T: 16–490T: 9–489T: 1–486

5oglA

(713 AA)

[ ]

2E-165.7E-144.5E-161.1E-15
Q: 11–455Q: 2–475Q: 1–426Q: 4–462
T: 1–432T: 18–343T: 12–433T: 1–434

6p25A/6p2rA

(817 AA)

[ ]

5.7E-147.8E-136.3E-141.8E-13
Q: 1–453Q: 2–473Q: 1–257Q: 1–303
T: 26–703T: 53–705T: 45–290T: 32–304

7bvfA

(1102 AA)

[ ]

1.9e-121.2e-101.1e-122.2e-11
Q: 33–456Q: 7–473Q: 13–426Q: 26–462
T: 263–631T: 263–631T: 263–633T: 263–630

6sniX/6snhX

(562 AA)

[ ]

1.5E-101.3E-081.8E-111.2E-09
Q: 1–416Q: 1–434Q: 1–388Q: 1–424
T: 30–411T: 54–411T: 49–411T: 35–411

The eight essentially full-length hits with best E-values and sequence coverage > 90% are tabulated: 5ezm, crystal structure of ArnT from Cupriavidus metallidurans in the apo state [ 58 ], 5f15 is the same as 5ezm but with undecaprenyl phosphate as analogue for a lipid-linked sugar substrate; 6s7t, cryo-EM structure of human oligosaccharyltransferase complex OST-B [ 59 ]; 6s7o, cryo-EM structure of human oligosaccharyltransferase complex OST-A [ 65 ]; 6ezn, cryo-EM structure of the yeast oligosaccharyltransferase (OST) complex [ 74 ]; 3waj, crystal structure of the Archaeoglobus fulgidus oligosaccharyltransferase (O29867_ARCFU) complex with Zn and sulfate [ 75 ]; 5ogl, structure of bacterial oligosaccharyltransferase PglB in complex with an acceptor peptide and an lipid-linked oligosaccharide analogue [ 60 ]; 6p25/6p2r, structure of Saccharomyces cerevisiae protein O-mannosyltransferase Pmt1-Pmt2 complex bound to the sugar donor and a peptide acceptor/without peptide ligand [ 59 ]; 7bvf_A, Cryo-EM structure of Mycobacterium tuberculosis arabinosyltransferase EmbA-EmbB-AcpM2 in complex with ethambutol [ 76 ]. We added also 6sni/6snh (cryo-EM structure of nanodisc reconstituted yeast ALG6 in complex with 6AG9 Fab or with Dol25-P-Glc [ 77 ]) because of the much shorter template length. For each query and each PDB structure (listed as PDB ID), we provide the E-value and the sequence ranges hit in the query (Q) and in the template (T; we also provide the length of the template in parentheses below the PDB identifier). The uppercase letter behind the PDB identifier denotes the relevant chain

Proteins with known structure discovered in these searches belong to the group of well-studied membrane-standing arabinosyl-, oligosaccharyl- or mannosyltransferases. Their annotated enzymatic domain is fully part of the alignment. Given the full-length coverage of the N-TMTCs’ sequences queried against the PDB, there is no doubt that N-TMTCs and the annotated enzymatic domains of sugar transferases detected share a common fold and have a similar 3D structure.

For all N-TMTCs, the sequence of the bacterial aminoarabinose transferase ArnT corresponding to structures 5ezm/5f15 [ 58 ] is the most similar homologue with an almost gapless alignment (with some exception for the N-terminal region of the loop between TM7 and TM8). The alignments of N-TMTCs generated by HHpred cover the first 11 of the 13 N-terminal TMs in 5ezm/5f15, nicely supporting the membrane topology consideration in the previous section (to note, TM region TM4 is missing and TM5/6 are annotated as a single large TM both in the PDB entry 5ezm and in the Uniprot entry Q1LDT6). As a result of the structural similarity, we can conclude that there are five loops between TM regions that form the structure in the ER lumen (see Fig. ​ Fig.2): 2 ): (i) two long loops EL1 (between TM1/TM2) and EL4 (between TM7/TM8; both loops contain helical segments) as well as (ii) three short loops EL2 (between TM3/TM4), EL3 (between TM5/TM6) and EL5 (between TM9/TM10). In 5ezm/5f15 (as in other sugar transferases of this type), there are two substrate binding cavities that communicate via a channel limited, on one side, by the TMs in the membrane and, at the other side, by the long loop connecting TM7 and TM8 (i.e., EL4 in the case of TMTCs). One binding region is formed by the segments homologous to EL1, EL2 and EL4 and accommodates the sugar acceptor substrate. The other site (built by EL1 and mainly by EL4) provides for interaction with a lipid-linked carbohydrate (LLC; the sugar donor, e.g., a dolichyl phosphate or pyrophosphate with attached sugar/oligosaccharide moiety). In the zone of contact of the two substrates, a divalent metal ion important for catalysis is coordinated by amino acid residues of the transferase. Despite the vast differences in sequences and possible ligands, homology considerations suggest that the TMTCs are constructed following the same general architecture.

Most importantly, we see at the level of sequence comparison (even without any structural modelling) that some critical motifs strongly conserved among the TMTCs have a structural and/or functional equivalent (e.g., in ligand binding) in the 3D structures of enzymes found. The strictly conserved DD motif in the loop between TM1 and TM2 (e.g., D52/D53 in N-TMTC1) aligns with the known active site in several sugar transferases (e.g., D55/E56 in 5ezm_A, D77/E78 in 6p25_A or D281/D282 in 7bvf_A). All the sugar transferases found in our HHpred homology search have at least an aspartate that coincides with the first aspartate in this motif. This residue is described as binding to the polar group of the sugar acceptor and/or a divalent metal ion (e.g., for 5ezm/5f15 [ 58 ], 5ogl [ 60 ], 6s7t/6s7o [ 65 ] or 6sni/6snh [ 77 ]). Thus, these positions are absolutely critical for enzymatic catalysis since any residue substitution leads to loss of function. For example in 6p25/6p2r [ 59 ], E78 forms a salt bridge with R138 making D77 sticking out towards the cavity where it binds to the sugar acceptor substrate. Any replacement of D77/E78 abolishes enzyme function [ 59 , 78 ].

In 5ezm/5f15, D158 (in EL2, N-terminal to TM4) interacts with the acceptor substrate and also forms a salt bridge with K203 (in EL3, C-terminal to TM5). The homologous residues are conserved in TMTCs (e.g., D169 and K219 in N-TMTC1) and, thus, are predicted to also play a role in ligand binding.

An arginine in the loop EL5 between TM9 and T10 close to the N-terminus of TM10 and strictly conserved among TMTCs (e.g., R404 in TMTC1 as part of the conserved sequence AERV) followed by a hydrophobic stretch of residues (from TM10) is also seen in sugar transferase structures (R459 in 6s7t [ 65 ], R405 in 6s7o [ 65 ], R404 in 6ezn [ 74 ], R426 in 3waj [ 75 , 79 ], and R375 in 5ogl [ 60 ]). In all these known structures, this arginine is described as an interaction partner of the LLC’s phosphate group whereas the lipid part of the LLC is accommodated within a hydrophobic groove formed mainly by TM6 and TM7.

The sequence SHKSYRP (with H89/K90 in TMTC1) in EL1 is well conserved among TMTCs (close to the N-terminal end of second helix in EL1). At the same time, K85 in the 5ezm/5f15 sequence at a homologous position is known to interact with the LLC’s phosphate. Thus, it is reasonable to assume that one of the positively charged residues in TMTCs (e.g., H89 or K90 in TMTC1) has a similar role. This suggestions is supported by the known mutant phenotype in human TMTC3 (the mutation His67Asp introduces a charge swap and leads to cobblestone lissencephaly [ 19 ]; H67 is the position in TMTC3 homologous to H89 in TMTC1).

The limits of a purely sequence-analytic approach can be illustrated with the case of the DW motif conserved among all TMTCs in EL4 (e.g., D330/W331 in N-TMTC1) at the C-terminal end of the helix parallel to the ER membrane. It is problematic to identify the function of an equivalent motif in homologous 3D structures, even in those with a hit to DUF1736. For example, the apparently homologous sequence position R270/Y271 in 5ezm/5f15 are at the edge of a structurally unresolved loop region. In 6s7t, residues E405/H406 seem the closest to positions homologous to the TMTCs’ DW motif. E405 is directed towards R214 (a residue in the loop homologous to EL2) [ 65 ]. Thus, the function of the conserved DW motif in TMTCs (as well as of several others) cannot be unambiguously understood due to such comparisons. Interestingly, a DW motif has been described as critical for subunit interaction in pyruvate dehydrogenase kinase 2 [ 80 ].

Thus, this sequence-analytic comparison of TMTCs with known homologous 3D structures shows that a number of conserved sequence motifs can be understood in the context of ligand binding. TMTCs appear to incorporate divalent metal ions for catalysis and LLCs as donors for a sugar moiety. Given the experimental finding of TMTCs being part of a new O-mannosylation pathway [ 26 ], the LLC applicable here is dolichyl-phospho-mannose (DPM), the universal donor of mannosyl-residues in higher eukaryotes.

TMTCs are homologous to a variety of sequence families of membrane-bound sugar transferases

When applying HHPred with N-TMTCs as input against the Pfam library of sequence domain family models, a large variety of annotated entries besides many domains of unknown function are hit with, beyond doubt, statistically significant E-values (E-value< 1.e-5, see Table  2 and Additional file 3 ).

HHPred search with N-terminal part of four human TMTCs against Pfam-A_v33.1

Pfam domainTMTC1
(1–456)
TMTC2
(1–475)
TMTC3
(1–426)
TMTC4
(1–462)
Glyco_transf_22 (PF03901, 388 AA)2.1E-201E-181.5E-206.4E-19
Q: 29–456Q: 3–475Q: 9–426Q: 22–462
T: 1–352T: 1–351T: 2–350T: 1–350

STT3

(PF02516, 458 AA)

5.5E-192.1E-171.8E-199.5E-18
Q: 26–456Q: 1–473Q: 5–423Q: 19–459
T: 3–406T: 4–400T: 3–401T: 3–401
PTPS_related (PF10131, 616 AA)1.4E-159.7E-142.5E-164.3E-15
Q: 89–456Q: 62–475Q: 67–425Q: 81–462
T: 1–308T: 1–308T: 1–307T:1–308

PMT

(PF02366, 247 AA)

2.3E-141.3E-131.5E-142.1E-13
Q: 30–285Q: 3–248Q: 9–259Q: 23–293
T: 2–242T: 1–242T: 2–242T: 2–242
Mannosyl_trans2 (PIG-V) (PF04188, 432 AA)6E-141.9E-123.8E-141.4E-12
Q: 51–451Q: 25–470Q: 30–426Q: 44–462
T: 60–425T: 60–425T: 60–429T: 60–427

Dpy19

(PF10034, 651 AA)

8.4E-131.8E-124.4.E-133E-12
Q: 46–45520–47427–42439–460
30–50230–503 (651)32–499 (651)30–499 (651)

AftA_N

(PF12250, 432 AA)

3.6E-123.7E-113.1E-131.6E-11
T: 27–446T: 3–465T: 7–399T: 20–435
Q: 76–430Q: 78–431Q: 76–402Q: 75–402

PMT_2

(PF13231, 159 AA)

3.7E-131.1E-116.3E-131.6E-12
Q: 91–276Q: 64–234Q: 69–250Q: 83–284
T: 1–156T: 1–156T: 1–156T: 1–159
Arabinose_trans (PF04602, 471 AA)6.3E-115.9E-091.1E-101.6E-10
Q: 34–456Q: 8–468Q: 13–426Q: 27–462
T: 51–428T: 51–423T: 51–430T: 51–427

PIG-U

(PF06728, 363 AA)

9.8E-117.8E-092.1E-105E-09
Q: 47–456Q: 6–475Q: 14–423Q: 45–462
T: 30–349T: 1–350T: 1–345T: 35–349
Mannosyl_trans4 (PF15971, 163 AA)9.4E-111E-096.2E-113.9E-10
Q: 81–276Q: 59–234Q: 59–250Q: 78–285
T: 1–162T: 6–161T: 1–162T: 6–162
Glucos_trans_II (PF14264, 312 AA)6.4E-072.8E-067.6E-085.6E-07
Q: 45–413Q: 19–431Q: 24–385Q: 38–421
T: 5–310T: 5–310T: 5–310T: 5–310

GT87

(PF09594, 251 AA)

3.5E-076.4E-061.1E-061.5E-06
Q: 91–389Q: 64–406Q: 68–360Q: 82–396
T: 2–251T: 2–249T: 1–248T: 1–248

The functionally annotated hits with best E-values are listed: PF03901, Alg9-like mannosyltransferase family; PF02516, Oligosaccharyl transferase STT3 subunit; PF10131, 6-pyruvoyl-tetrahydropterin synthase related domain, function unknown; PF02366, Dolichyl-phosphate-mannose-protein mannosyltransferase; PF04188, Mannosyltransferase (PIG-V); PF10034, Q-cell neuroblast polarisation, function unknown; PF12250, Arabinofuranosyltransferase N terminal domain; PF13231, Dolichyl-phosphate-mannose-protein mannosyltransferase; PF04602, Mycobacterial cell wall arabinan synthesis protein; PF06728, GPI transamidase subunit PIG-U; PF15971, Dolichyl-phosphate-mannose mannosyltransferase; PF14264, Glucosyl transferase Gtr II; PF09594, Glycosyltransferase family 87. For each query and each Pfam entry (listed as Pfam entry name and ID), we provide the E-value and the sequence ranges hit in the query (Q) and in the template (T; we also provide the length of the template in parentheses below the Pfam model name)

Most of the domains found belong to the GT-C clan (CL0111) of glycosyltransferases (out of 19 known GT-C members, nine were detected: Glyco_transf_22, STT3, PTPS_related, PMT, Mannosyl_trans2, PMT_2, Arabinose_trans, PIG-U, GT87). Most informative are the sequence homologies with Glyco_transf_22 (PF03901) and STT3 (PF02516) because the E-value is < 1.e-18 and alignment of the Pfam domains and the N-TMTCs cover both query and template almost completely (coverage > 95%). Certain super-conserved residues in the sequence family alignments of both Pfam families are also conserved among the TMTCs. This includes the active site DD motif in EL1 (e.g., D52/D53 in N-TMTC1) and the arginine in front of TM10 (e.g., R404 in TMTC1) that are characteristic for both Pfam domains.

The homology with other groups of dolichyl-phosphate-mannose-dependent mannosyltransferases (Mannosyl_trans4, PF15971), glucosyl transferases GtrII (Glucos_trans_II, PF14264) and arabinofuranosyltransferase N-terminal domain (AftA_N, PF12250) not directly linked to the GT-C clan fits into the same general functional prediction for TMTCs as sugar transferases and having a similar 3D structure.

The HHPRED search results are confirmed by iterative PSI-BLAST [ 32 ] runs with standard parametrization and human TMTC sequences as input. They deliver plentiful hits within the GT-C clan and beyond (results not shown). The diversity of significant homology hits constitutes a problem for function assignment of TMTCs beyond the general prediction as GT-C/PMT-like sugar transferases. It needs to be emphasized that the GT-C clan is a very diverse sequence superfamily comprising membrane-bound sugar transferases with a large variety of different specific activities and substrate types (including the transfer of arabinose, mannose, glucose or oligosaccharides among others).

We find also other proteins including even enzymatically completely inactive ones such as PIG-U (see reference [ 81 ] for discussion of PIG-U’s function). Interestingly, the profile build on the basis of our grand alignment of TMTCs is linked by HHPred to the domain BindGPILA [ 81 ] with E-value ~ 0.03 (calculated at the background of all Pfam models). To note, this domain model is derived from homologous sequence segments with 10 TMs and intermittent loops extracted from proteins in the glycosylphosphatidylinositol (GPI) lipid anchor pathway PIG-B, PIG-M, PIG-U, PIG-V, PIG-W and PIG-Z [ 81 ]. PIG-W is an acetyltransferase for the GPI lipid anchor, PIG-U is not an enzyme at all but the remaining four (PIG-B, PIG-M, PIG-V and PIG-Z) are mannosyltransferases. All of them are united by the ability to bind phospho-lipid linked sugar/carbohydrate moieties.

Thus, the mere homology of TMTCs to the GT-C group of sequences by itself is only informative with regard to fold coincidence, to structural similarity and to a general level of functional classification. Yet, the conservation of residues known to be important for catalysis and substrate binding as detailed in the sequence analysis above indicates that TMTCs are actually enzymatically active. As we see in the 3D structure modelling exercise below, many additional conserved sequence motifs can be rationalized due to interactions with ligands and substrate molecules.

Insights from the structural modelling of human TMTCs by homology to membrane-bound sugar transferases with known 3D structural arrangements

We attempted to create 3D structural models of all four TMTCs together with a divalent metal ion and DPM with the goal to explore whether observed sequence motifs that are conserved between TMTCs and sugar-transferases of known 3D structure come spatially together for interaction with the ligands.

HHpred scored the aminoarabinose transferase structures ArnTCm (PDB IDs: 5ezm and 5f15, chain A [ 58 ]) as by far the best hit for all human TMTCs (see Table ​ Table1) 1 ) and also for five other organisms including Bos taurus, Gallus gallus, Danio rerio, Xenopus laevis and Drosophila melanogaster (results not shown). Therefore, this X-ray crystal structure was used as a template to build 3D models of TMTC1 (XP_016875493.1), TMTC2 (Q8N394), TMTC3 (Q6ZXV5) and TMTC4 (Q5T4D3) using the functions automodel and loop refine in Modeller (version 9.4) [ 35 ]. The overall structure of 5ezm (apo ArnTCm, resolution 2.70 Å) / 5f15 (UndP-bound ArnTCm, resolution 3.20 Å) [ 58 ] consists of (i) an N-terminal membrane-embedded region and (ii) a periplasmic domain (PD). For this work, only the first segment is of interest. It involves 13 TM helices and interconnecting loops including three juxtamembrane helices (JM1, JM2 and JM3). JM1 and JM2 form the first periplasmic loop between TM1 and TM2 while JM3 leads into a partially disordered flexible periplasmic loop (PL4 being homologous to EL4 in TMTCs) between TM7 and TM8.

In this study, only the membrane-embedded domain of TMTCs including the juxtamembrane helices were modelled using the most N-terminal regions of the templates 5ezm and 5f15 (the 11 TM segments together with JM1 and JM2 following 5ezm while JM3 was molded after 5f15). The major hurdles to generate the 3D structure of TMTCs by homology modelling are (i) the low percent identity (< 15%) with sequences of the template crystal structures (Table  3 ) and (ii) several overly long loops between TM regions without equivalent in the structure templates. As we want to understand structural detail at the lumenal side, cytoplasmic loops are not that critical but the lumenal ones are. The loop sequence segments include (i) the cytoplasmic loop between TM2-TM3 (residues 136–146) in TMTC4, (ii) the cytoplasmic loop between TM6-TM7 in all TMTCs and (iii) the lumenal loop TM9-TM10 in all TMTCs. Furthermore, the template 5ezm/5f15 does not account for a loop extension at the N-terminal side of the domain of unknown function, DUF1736 (PF08409), between TM7-TM8 for all TMTCs. Moreover, we note that TMTC2 has another unusually longer cytoplasmic loop between TM8-TM9 (residues 337–392) and, therefore, in the absence of any template, residues 337–392 were not modelled. We describe the alignment with the 5ezm/5f15 template, the regions modelled for each TMTC proteins and issues with the overly long loops in Table ​ Table3 3 and in the annotated alignment in Additional File  4 – Supplementary Figure 1.

Modelling the 3D structures of TMTCs

TMTC1TMTC2TMTC3TMTC4

9.4%10.6%9.5%11.3%
23–4561–336 & 393–4744–42817–464
240–257207–220209–231242–262
393–406411–419*365–373401–409

284–358 (321–335)247–321 (284–298)258–331 (294–308)292–366 (329–343)

The table provides the sequence identities of template 5EZM/5F15 with TMTCs, the range of the modelled regions, the longer loops between TM6-TM7 and TM8-TM9 compared with the templates, and location of DUF1736 along with JM3 (*residues 365–369 continue to be helical with TM9). TMTC2 has another unusual, longer cytoplasmic loop between TM8-TM9 (residues 337–392) which is not modelled in the absence of any template

As we expect that certain long loops, especially those that have no equivalent in the 5ezm/5f15 structure, will not get reconstructed well, the DOPE model scoring system provided by Modeller might not be such a good choice for selecting among various model instances. We have validated our model instances based on the TM-align scores [ 82 ]. A TM-score between 0 and 0.3 suggests random structural similarity while a TM-score greater than 0.5 and less than 1.0 suggests two structures having the same fold. The TM-align scores for TMTC1, TMTC2, TMTC3 and TMTC4 (when compared with 5ezm) are 0.93441, 0.72261, 0.91499, and 0.92104 respectively.

The resulting 3D structure models (see Fig.  3 ) were used to place a divalent metal ion (following 5ezm for initial positioning) and a DPM moiety (using crystal-bound ligand UndP in 5f15 for initial posing as reference position). We applied Zn + 2 parametrization for the ion in this study although there is no clarity about the exact nature of the divalent metal ion from experiment. The crystallographic evidence speaks for zinc in 5ezm [ 58 ]; yet, Mn 2+ is the likely ion in the case of 5ogl [ 60 ], several other reports such as the one for 6s7t [ 65 ] remain silent about the nature of the ion other than emphasizing an electronic density consistent with a divalent metal ion. To emphasize, we do not think that the exact parametrization of the ion (beyond carrying two positive charges) is critical for the outcome of this modelling study.

An external file that holds a picture, illustration, etc.
Object name is 13062_2021_291_Fig3_HTML.jpg

Structure models of TMTC1/2/3/4 with ligands. The cartoon representation of model TMTC1/2/3/4 (from top to bottom) with docked DPM is shown in side- (left column) and top-view (middle column). Close-up (right column) of the binding pocket of TMTCs with docked DPM (cyan color sticks) and with important residues (HKSY residues of the conserved SHKSYRP motif M2 in EL1; K and E from motif M4 in EL3) presented in yellow color sticks; the divalent metal ion (modelled as zinc) is shown in gray color

3D structure modelling operations including ligands were implemented with Schrodinger suite [ 36 ]. An induced fit procedure following established protocols [ 36 – 42 ] was applied. In brief, the Schrodinger programs “Protein Preparation Wizard” and “LigPrep” were utilized for preparing the TMTC models and the DPM. With “Glide-SP” and “Prime”, multiple poses of DPM were generated and optimized in multi-step energy minimizations (with the OPLS parameter set and a surface Generalized Born implicit solvent model) that included some stages with softened potentials and side chains mutated to alanine. The procedure was completed with a minimization that allowed all residues within 5 Å of DPM (including their backbone and side-chain) and ligand DPM itself to be relaxed. The complexes were ranked by Prime energy (molecular mechanics energy plus solvation) and those within 30 kcal/mol of the minimum energy structure were passed through for a final round of Glide docking and scoring with GlideScore. The final structures for each of the TMTCs together with the ligands are provided with their atomic coordinates (Additional File  5 ).

As the most important outcome of the modelling effort, visual inspection of the four model structures show that, for all TMTCs, the resulting structures show consistently that seven conserved sequence motifs M1-M7 as listed in Table  4 come spatially together at the lumenal side of the TMTCs, form part of the surface of the protein structure that is homologous to the two substrate/ligand binding sites in 5ezm/5f15. They group closely around the DPM moiety and the divalent ion creating a dome region (see Fig.  4 for the case of TMTC1). We find that residues in motifs M4 and M5 are observed for coordinating the divalent metal ions. M2 and M3 are largely engaged in mannose interactions, M6 tends to contact with the dolichyl tail. Motifs M4, M5 and M7 are important for interaction with the phosphate in DPM. Thus, the observed sequence conservation can be rationalized in terms of evolutionary conserved function.

Several conserved sequence motifs in TMTCs are related to DPM binding and divalent metal ion coordination

MotifResiduesTMTC1TMTC2TMTC3TMTC4

M1 (red)

DD in EL1

D52263145
D53273246

M2 (orange)

SHKSYRP in EL1

mannose

S88 61 66 80
H89 62 67 81
K90 63 68 82
S91 64 69 83
Y92 6570 84
R93 667185
P94677286

M3 (yellow)

RxD in EL2

R167139 143172
D169141 145 174

M4 (green)

KE(T/Q) xxT in EL3

K219 186 188 221
T/Q221(T) 188(Q) 190(Q) 223(Q)
T224 191 193 226

M5 (blue)

DW in EL4

W331294 304 339

M6 (violet)

PxxP in TM9

P386 404 358 394
P389 407 361 397

M7 (pink)

ERxxY in EL5

E403 421 375 411
R404 422 376 412
Y407 425 379415

Conserved residues present in the vicinity of the ligand dolichyl-phosphate-mannose (DPM) are part of seven motifs M1-M7 in the TMTC family protein sequences. For each motif, the actual sequence, the location (loop number or TM number), loop coloring in Fig. ​ Fig.4 4 and the residue numbers in TMTC1/2/3/4 respectively are listed. If at least one atom of the residue is within 5 Å, 6 Å or 7 Å of any atom of DPM, the respective residue is marked with the corresponding subscript “A”, “B” or “C”. In bold, we indicate residues in M4 and M5 observed for coordinating the divalent metal ions. We find motifs M2 and M3 largely involved in mannose interactions, M6 provides for the dolichyl tail, and M4, M5 and M7 are important for interaction with the phosphate

An external file that holds a picture, illustration, etc.
Object name is 13062_2021_291_Fig4_HTML.jpg

Sequence motifs M1-M7 come spatially together in model structures of TMTCs. We illustrate the spatial localization of sequence motifs M1 (red), M2 (orange), M3 (yellow), M4 (green), M5 (blue), M6 (violet) and M7 (pink, all shown in ball mode) at the background of the structural cartoon of the whole protein. DPM is presented as blackish sticks, the divalent metal ion is represented as reddish sphere. We show the case of TMTC1; the figures for the other TMTCs look very similar. To note, motif M2 in this figure is extended to the conserved region represented by SHKSYRPLCVTLTSFRLN in TMTC1 (88–103 in EL1)

Further, several close contacts between the DPM ligand, the metal ion and TMTC residues were observed (to note, we did not enforce any specific residue contacts during the induced fit docking procedure). Given some sequence diversity among TMTCs and also the large number of degrees of freedom in the modelling process, it is not surprising that not all contacts are found in all models. Yet, a common subset of those was detected in each of the TMTC1, TMTC2, TMTC3, and TMTC4 model structures (see Table ​ Table4) 4 ) and some contacts repeat patterns seen in homologous crystal structures:

  • (i) The phosphate functional group of DPM interacts with the divalent metal ion. In addition, the metal binds to the glutamate residue in the conserved KET(Q) xxT motif in EL3 (e.g., E220 of TMTC1) and to aspartate residue of the conserved DW motif (e.g., D330 in TMTC1) in EL4. To note, H267 (in the motif H265-E266-H267 where the glutamate is homologous to D330 in TMTC1) interacts with the divalent metal ion held between JM1 and EL4 in 5ezm [ 58 ].
  • (ii) The phosphate group of DPM also forms a salt bridge with the lysine residue of the conserved KET(Q) xxT motif in EL3 (e.g., K219 in TMTC1).
  • (iii) The mannose moiety interacts with residues H-K-S-Y within the conserved SHKSYRP motif M2 in EL1 (e.g., S80, H89, K90 and S91 residues in TMTC1, Fig. ​ Fig.3 3 ).
  • (iv) The conserved stretch in EL1 represented by SHKSYRPLCVLTSFRLN in TMTC1 (it includes motif M2) forms the dome region of the DPM binding pocket in all 4 TMTCs. The dolichyl lipid chain of DPM occupies the cavity that is provided by hydrophobic residues of TM6, TM7 and TM9.

The structural models of human TMTCs can only be considered preliminary in many details at this stage since

  • important ingredients such as the protein substrate and possibly important interacting partners are missing,
  • sequence identity with the target structure is low (~ 10% in the manually edited alignments used for modelling, Table ​ Table3 3 ),
  • there are loop extensions not found in the structural template, and
  • the TMTCs are modelled without the C-terminal TPR domain.

The average accuracy of C-alpha atom positioning in homology modelling above 30% sequence identity is estimated 2 Å [ 83 , 84 ]; hence, the error is expected to be higher for certain regions in our model structures, especially in loop regions without equivalent in the template. On the other hand, the known crystal structures (having very moderate crystallographic resolutions around 3 Å) do not resemble the complete protein complex including the correctness of certain groups of amino acid chains, some inter-TM loops, substrates and ligands needed for catalysis either.

Despite these restrictions, we see consistent features emerging from the modelling of various TMTCs, namely the arrangement of TM regions in the membrane as well as of the loops and segments that form the binding site for the lipid-linked sugar and the divalent metal ion; essentially, the major part of the structure located in the ER lumen appears functionally plausible after the conserved sequence segments got spatially united as a result of the 3D reconstruction.

Thus, it makes sense to analyze also contacts between the DPM moiety, the metal ion and TMTC residues seen only in a few of the TMTC models. In this way, we will get a more complete picture of the binding cavity and can enlarge the list of potentially relevant residues for interaction with the ligands:

  • (i) We found the aspartate from motif M3 in the vicinity of the mannose in TMTC2 (D141) and TMTC3 (E145). The homologous residue D158 in 5f15 [ 58 ] is also seen to interact with the arabinose moiety.
  • (ii) K203 in 5f15 [ 58 ] forms a salt bridge to the arabinose moiety. A similar close contact to the sugar is seen by homologous lysine residues in motif M4 for TMTC2 (K186), TMTC3 (K188) and TMTC4 (K221).
  • (iii) The motif M7 arginine in TMTC2 (R422) forms a hydrogen bond with the phosphate. This interaction resembles the contact between several homologous arginine residues (R459 in 6s7t [ 65 ], R405 in 6s7o [ 65 ], R404 in 6ezn [ 74 ], R426 in 3waj [ 75 , 79 ], and R375 in 5ogl [ 60 ]) and the phosphates from the respective LLCs in those X-ray 3D structures. Similarly, the M7 tyrosine is observed close to the phosphate in TMTC2 (Y425) and TMTC4 (Y415) as Y345 in 5f15 [ 58 ].
  • (iv) Residues E84/K85 in 5ezm [ 58 ] do interact with the metal ion in the absence of a LLC molecule. We see the homologous residues HK in motif M2 also interacting with a ligand (but with the sugar moiety) in our TMTC models.

Despite the wealth of sequence-analytic findings available for TMTCs, the systematic analysis of their sequences and of related biomolecular data for the purpose of assigning the biological function of TMTCs has never been performed before. Several roadblocks had to be overcome. First, there are issues with sequence accuracy as, for some TMTCs, several versions of protein sequences are available in databases, some of which lack sequence pieces essential for TMTC function as this study has revealed. Second, the complex nature [ 66 ] of the TM regions sprinkled with polar residues/prolines/glycines makes their accurate prediction in the TMTC sequences difficult. This seriously hampers function discovery since localizing certain loops at the correct side of the membrane might be impossible with errors in membrane topology. Third, just the fact of finding sequence similarity with a large number of sugar transferases is helpful to establish the homology relationship but provides little guidance for biological follow-up work aimed at zooming into the exact molecular and cellular functions of TMTCs, for example with regard to actual catalytic capacity, substrate specificity and ligands bound.

This work has made significant steps forward in understanding 3D structure and biological function of the membrane-embedded domains covering the N-terminal halves of TMTC1, TMTC2, TMTC3 and TMTC4 sequences. First, we determined the exact membrane topology using sequence-analytic, phylogenetic and available experimental data. The assumption of conserved membrane topology for evolutionarily conserved molecular function was key to interpret TM prediction results for N-TMTCs in a unified manner. The finally determined membrane topology including 11 TMs nicely complies with all known constraints. The C-terminal globular TPR domain is located in the ER lumen together with the critical for function conserved sequence motifs in the loops between TM regions. The homologous sequence segments in the known 3D structures 5ezm/5f15 corresponding with the luminal loops in TMTCs have the same membrane topology. We can further conclude that TMTC sequences in the database that cannot fit to this topology are most likely erroneous.

Whereas the complex nature of TM regions in TMTCs makes TM prediction difficult, it supports establishing gene homology via searches for significant sequence similarity [ 66 , 70 ]. The evidence certifying the homology of N-TMTCs with GT-C/PMT-class and other related sugar transferases is overwhelming; thus, TMTCs must have the same overall fold and resemble similar tertiary structure. Despite the huge evolutionary distance from bacteria to human representatives in this homology group, higher eukaryote TMTCs share strongly conserved sequence motifs with GT-C/PMT-class enzyme sequences. Even at the pure sequence-analytic level, we can explain a few of these conserved sites as required for catalysis or for ligand binding. Given the close relationship with ArnT from Cupriavidus metallidurans (the structure of which is known: 5ezm/5f15), we suggest that these ligands include a divalent metal ion and a LLC molecule. Since TMTCs are part of an O-mannosylation pathway, we conclude that this LLC is DPM.

3D-structural modelling of N-TMTCs further enhances the association of conserved sequence motifs with ligand binding. Seven conserved sequence motifs from various parts of the protein sequence (including those seen already at the level of just sequence comparison) come spatially together to form the surface of binding sites for the mannosyl residue, the phosphate group and the dolichyl tail of DPM as well as the divalent metal ion; thus, their evolutionary conservation can be rationalized as maintaining the ability to position these two ligands for catalysis. Notably, this spatial co-localization of peptide stretches corresponding to the conserved motifs is sufficiently macroscopic to be a reliable result not affected by the accuracy of the homology procedure applied here.

In addition, we derive, as a result of this homology-supported structural modelling, a further expanded list of residues taken from the set of conserved motifs that are potentially interacting with the divalent metal ion and the DPM ligand. This list comprises those critical residues previously found with combined phylogenetic arguments (sequence conservation among TMTCs and similarity with sequences of structurally and functionally characterized sugar transferases) as a subset. Thus, we can relate certain residues strictly conserved among the TMTC sequences with functions in catalysis and ligand binding. This work also clarified the nature of the DUF1736 sequence segment in TMTCs, actually a loop between TM7 and TM8 the accurate positioning of several of its functional residues is critical for catalysis and binding of ligands, especially the lipid-linked sugar moiety.

Notably, we have already established the homology of TMTCs with GT-C/PMT-class sugar transferases when we first analysed their sequences for the first time in 2012; yet, a substrate and biological context assignment as well as 3D structural modelling were not possible. With HHpred [ 33 ], significant sequence similarity with DPM-dependent mannosyltransferases (PMTs, PF02366) was detected. With RPS-BLAST [ 85 , 86 ], we found the link to ArnT-like arabinose transferases (COG1807). Their respective 3D structures were not known during that time [ 58 ].

The density of hints derived from sequence analysis, phylogenetic comparisons, homology studies and structural modelling leaves no doubt that the TMTCs have enzymatic activity and perform sugar moiety transferase functions in their biological context. Thus, the O-mannosyl-transferase sought in the recently discovered new O-mannosylation pathway (via combinations of TMTC knock-outs) that selectively processes cadherin-like targets and that the TMTCs are members of [ 26 ], are actually the TMTCs.

Finding the real substrates of the various human TMTCs and rationalising the function of their glycosylation are important questions from the view-point of biological science. Additionally, this topic has a critical medical dimension as several mutations of TMTCs are compatible with survival but severely disable the affected patients in various ways due to the pleiotropic nature of their molecular and cellular functions. Laudably, first steps in this direction have been done. It can be concluded that various cadherins/proto-cadherins found as substrates for the new O-mannosylation pathway are protein substrates for O-mannosylation by TMTCs [ 25 , 26 ].

BLAST/PSIBLAST [ 32 ] searches reveal TMTC proteins are present in a wide range of animals but apparently not in fungi and plants (details not shown). Interestingly, essentially full-length homologous sequences (including the sugar transferase followed by TPR segments) are also found in many, typically not yet well characterized prokaryotes besides hits in lower eukaryotes such as oomycetes and choanoflagellates. One example is protein AMJ42_05695 (from Deltaproteobacteria bacterium DG_8) that is found by a BLAST search with human TMTC3 (24% sequence identity, E-value=3.e-47, alignment of query positions 12–698 against positions 46–774 from target). Human curiosity will not be satisfied until the diversity of their organic chemistry, the related biomolecular mechanisms and the cellular phenotypes will be understood.

Acknowledgements

This work has been supported by the Biomedical Research Council of the Agency for Science, Technology and Research (A*STAR). We thank Shazib Pervaiz (National University Singapore) for drawing our attention to the sequence-based function prediction of human TMTC2 in 2012.

Abbreviations

AAAmino acid(s)
DPMDolichyl-phosphate-mannose
EREndoplasmic reticulum
GPIGlycosylphosphatidylinositol
PDBProtein Data Bank ( )
TMTransmembrane (region)
TMTCTransmembrane and tetratricopeptide repeat-containing
TPRTetratricopeptide repeat

Authors’ contributions

BE and FE initiated the project and designed the computational approaches. BE, SS, VS, QWT, and FE made the sequence-analytic computations and evaluated the results. VS and FLS initiated the structural modelling; the final models were created by SS and CKJ and evaluated by FE. BE and FE were the major contributors in writing the manuscript. All authors read and approved the final manuscript.

Authors’ information

VS and QWT had several months of student internships in the sequence analysis group of BE and FE at the Bioinformatics Institute Singapore (QWT in 2016 and VS in 2019).

There was no dedicated funding for this project. The writing of this article benefitted from the shutdown of other activities during the COVID-19 lock-down. General financial support from A*STAR is gratefully acknowledged. QWT received an A*STAR Graduate Academy (AGA) scholarship (AUS) for her university studies. The internship of VS was supported by a SIPGA grant from AGA.

Availability of data and materials

Ethics approval and consent to participate.

Not applicable.

Consent for publication

Competing interests.

The authors declare that they have no competing interests.

Publisher’s Note

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

Birgit Eisenhaber and Swati Sinha joint first authors.

Contributor Information

Birgit Eisenhaber, Email: gs.ude.rats-a.iib@etigrib .

Frank Eisenhaber, Email: gs.ude.rats-a.iib@eknarf .

Academia.edu no longer supports Internet Explorer.

To browse Academia.edu and the wider internet faster and more securely, please take a few seconds to  upgrade your browser .

Enter the email address you signed up with and we'll email you a reset link.

  • We're Hiring!
  • Help Center

paper cover thumbnail

COMMENTS

  1. Lock-and-key model Definition and Examples

    Lock-and-key vs. Induced Fit Model. At present, two models attempt to explain enzyme-substrate specificity; one of which is the lock-and-key model, and the other is the Induced fit model.The lock and key model theory was first postulated by Emil Fischer in 1894.The lock-and-key enzyme action proposes the high specificity of enzymes.

  2. Lock and Key Model- Mode of Action of Enzymes

    Lock and Key Model. A German scientist, Emil Fischer postulated the lock and key model in 1894 to explain the enzyme's mode of action. Fischer's theory hypothesized that enzymes exhibit a high degree of specificity towards the substrate. This model assumes that the active site of the enzyme and the substrate fit perfectly into one another ...

  3. The Lock and Key Theory: Understanding Enzyme Specificity and Catalysis

    The Lock and Key Theory, postulated by Emil Fischer in 1894, is a cornerstone concept in biochemistry that elucidates the specificity of enzyme action. Enzymes are specialized proteins that act as catalysts to accelerate chemical reactions within biological systems. This theory analogizes the enzyme's active site to a lock and the compatible ...

  4. Molecular Docking: From Lock and Key to Combination Lock

    Figure 1. Illustration of 'Lock and Key' (top), Induced fit (middle) and Combination Lock (bottom) model of protein-ligand binding interaction. But, enzymes show conformational flexibility and, on that basis, Daniel Koshland proposed a modification to the 'lock and key' model. Koshland's suggestion was that active sites of enzymes are ...

  5. Lock-Key Model

    Lock-key model, or its modified version, the induced-fit model [21], explains catalysis by an enzyme with an easily accessible active site, while it is less appropriate for the enzymes with active sites buried in the protein core. The activity and specificity of such proteins is determined by not only the geometry and properties of the active ...

  6. Molecular Recognition: Lock-and-Key, Induced Fit, and ...

    The Induced Fit Model Builds upon the Lock-and-Key Hypothesis. This lock-and-key model served the biochemical community well for over 50 years. However, while this model adequately explained how substrates that are too large to fit within the confines of the active site would fail to act as substrates, it did not explain how small substrates, for instance water, often acted as non-substrates ...

  7. Molecular Recognition: Lock-and-Key, Induced Fit, and ...

    In 1894, Emil Fisher discovered that glycolytic enzymes are able to distinguish between sugar stereoisomers. Based upon that discovery, he formulated the lock-and-key hypothesis (Fischer 1894), which proposed that enzymes recognize their substrates just as a lock receives a key.That is, only in the case of exact geometric complementarity between the substrate (key) and enzyme (lock) is the ...

  8. The Central Role of Enzymes as Biological Catalysts

    The simplest model of enzyme-substrate interaction is the lock-and-key model, in which the substrate fits precisely into the active site (Figure 2.24). ... This example illustrates several features of enzymatic catalysis; the specificity of enzyme-substrate interactions, the positioning of different substrate molecules in the active site, and ...

  9. PDF Chapter 10 Specificity of Enzymes

    similar structure. The specificity of an enzyme with a substrate can be explained by "Lock and key" model. In this model, the lock and key correspond to the enzyme and the substrate, respectively, and only the correctly shaped key can fit into the key hole (active site). This theory is based on the "rigid enzyme" model

  10. Limitations and Extensions of the Lock-and-Key Principle: Differences

    1. Introduction. After Emil Fischer coined the lock-and-key picture for the reaction between enzymes and substrates [], it became a leading concept for the understanding of intermolecular interactions with proteins, and later for the rational design of drugs.With the advent of supramolecular chemistry the idea gained an enormous momentum, as chemists began to synthetize a large variety of host ...

  11. Introduction to enzymes and their applications

    'Lock and key' hypothesis of enzyme specificity. Harden and Young: 1901-3: Methods for the derivation of kinetic rate laws; principle of enzyme-substrate complex. ... Enzyme specificity is the absolute specificity of protein catalysts to identify and bind to only one or a few molecules. In this process the enzyme carries a defined ...

  12. Enzymes and the active site (article)

    Explore how enzymes work and how they shape the active site to catalyze biochemical reactions . Khan Academy offers a free, world-class education for anyone, anywhere.

  13. Key-lock hypothesis

    In protein: The role of the active site. …and enzyme, called the "key-lock" hypothesis, was proposed by German chemist Emil Fischer in 1899 and explains one of the most important features of enzymes, their specificity. In most of the enzymes studied thus far, a cleft, or indentation, into which the substrate fits is found at the active….

  14. Fundamentals of Enzymatic Processes

    Enzyme specificity is due to the way an enzyme interacts with the substrate molecule to form an enzyme-substrate complex (also called transition-state complex ). ... Lock and Key Theory states both the structure of enzyme and the substrate are rigid whereas Induced Fit Theory describes that the structure of enzyme is partially flexible. The ...

  15. How Emil Fischer was led to the lock and key concept for enzyme specificity

    How Emil Fischer was led to the lock and key concept for enzyme specificity Adv Carbohydr Chem Biochem. 1994:50:1-20. Authors R U Lemieux 1 , U Spohr. Affiliation 1 Department of Chemistry, University of Alberta, Edmonton, Canada. PMID: 7942253 No abstract available ...

  16. (PDF) A pilot screening of prevalence of atopic states and

    A pilot screening of prevalence of atopic states and opisthorchosis and their relationship in people of Tomsk Oblast

  17. PAPER OPEN ACCESS $VVHVVPHQWRI0HVR]RLF .DLQR ...

    According to recent data [2] permafrost was not only in the central and southern parts of West Siberia, but Quaternary ice cover embraced northern and northeastern parts of Kazakhstan; ice sheet

  18. Conserved sequence motifs in human TMTC1, TMTC2, TMTC3, and TMTC4, new

    Additional file 4 : Supplementary Figure 1.Alignment of TMTC sequences with those of the template structures 5ezm and 5f15 used for homology modelling. The file AF4-2020-10-modeller-alignment-TMTCs.pdf shows the alignment of the four human TMTC sequences with template structures 5ezm and 5f15 that was actually used for generating their 3D model with the Modeller suite version 9.4.

  19. Specificity of ethnogeny indigenous peoples by Central Siberia in the

    This article focuses on the processes of ethnic and cultural identification and self-identification, which the indigenous peoples of the North of Russia and Siberia, living in the Russian Federation, are currently going through. The post-Soviet