The Structure of Water

Learning objectives.

  • Describe the structure of a water molecule.
  • Describe the polarity of a water molecule.

How much of the Earth’s surface is covered with water?

Water is an extremely large component of many systems and is very important

In his well-known poem “The Rime of the Ancient Mariner,” Samuel Coleridge wrote “Water, water everywhere, Nor any drop to drink.”  The narrator was talking about being out on the ocean, but not having any water because he had killed an albatross (apparently bringing bad luck to everyone on the ship).  About 75% of the Earth’s surface is water.  The major constituent of the human body (over 60%) is water.  This simple molecule plays important roles in all kinds of processes.

Structure of Water

(delta -)

The water molecule, visualized three different ways: ball-and-stick model, space-filling model, and structural formula with partial charges.

The bent shape of the water molecule is critical because the polar O-H bonds do not cancel one another and the molecule as a whole is polar. Figure 2 below illustrates the net polarity of the water molecule.  The oxygen is the negative end of the molecule, while the area between the hydrogen atoms is the positive end of the molecule.

Water is a very polar molecule because the oxygen is very electronegative

Figure 2. Water is a polar molecule, as greater electron density is found around the more electronegative oxygen atom.

Polar molecules attract one another by dipole-dipole forces as the positive end of one molecule is attracted to the negative end of the nearby molecule.  In the case of water, the highly polar O-H bonds results in very little electron density around the hydrogen atoms.  Each hydrogen atom is strongly attracted to the lone-pair electrons on an adjacent oxygen atom.  These are called hydrogen bonds and are stronger than conventional dipole-dipole forces.

Hydrogen bonds are formed between electron-deficient hydrogen atoms and atoms with lone pairs

Figure 3. A hydrogen bond is the attraction between a lone pair of electrons on the oxygen atom of one molecule and the electron-deficient hydrogen atom of a nearby molecule.

Because each oxygen atom has two lone pairs, it can make hydrogen bonds to the hydrogen atoms of two separate other molecules.  The figure below shows the result – an approximately tetrahedral geometry around each oxygen atom consisting of two covalent bonds and two hydrogen bonds.

The geometry around each oxygen atom in water is tetrahedral because it has two covalent and two hydrogen bonds

Figure 4. As a result of two covalent bonds and two hydrogen bonds, the geometry around each oxygen atom is approximately tetrahedral.

  • Water is a molecular compound consisting of polar molecules that have a bent shape.
  • The oxygen atom acquires a partial negative charge while the hydrogen atom acquires a partial positive charge.

Use the link below to answer the following questions:

http://www.lsbu.ac.uk/water/molecule.html

  • What is the H-O-H bond angle?
  • How far is the center of each H atom from the center of the O atom?
  • What type of bond exists in a water molecule?
  • Which part of the molecule has a partial positive charge?
  • Which part of the molecule as a partial negative charge?
  • How do water molecules interact with one another?
  • electronegativity: A measure of the ability of an atom to attract the electrons when the atom is part of a compound.
  • polar bond:  A covalent bond in which electrons are shared unequally between two electrons.
  • Chemistry Concepts Intermediate. Authored by : Calbreath, Baxter, et al.. Provided by : CK12.org. Located at : http://www.ck12.org/book/CK-12-Chemistry-Concepts-Intermediate/ . License : CC BY-NC-SA: Attribution-NonCommercial-ShareAlike
  • Northern Bering Sea. Authored by : Courtesy of NOAA. Provided by : Wikimedia. Located at : http://commons.wikimedia.org/wiki/File:Bering_Sea_sunset_-_NOAA.jpg%20 . License : Public Domain: No Known Copyright
  • Water molecules. Authored by : Ben Mills (Wikimedia: Benjah-bmm27). Located at : http://commons.wikimedia.org/wiki/File:Water-3D-balls.png . License : Public Domain: No Known Copyright
  • water molecultes. Authored by : Ben Milles. Provided by : Wikimedia. Located at : https://commons.wikimedia.org/wiki/File:Water-3D-vdW.png . License : Public Domain: No Known Copyright
  • Wasser. Provided by : Wikimedia. Located at : https://commons.wikimedia.org/wiki/File:Wasser.png . License : Public Domain: No Known Copyright

Water Structure and Science (was www.lsbu.ac.uk/water)

This page forms the entrance to a website concerned with water's physical, chemical, and biological properties.

Table of Contents | Site Map

Liquid water is not a bit player in the theatre of life — it is the headline act

Many regard water (H 2 O) as a rather uninteresting substance because it is transparent, odorless, tasteless, and ubiquitous. It is the simplest compound of the two most common reactive elements in the Universe, consisting of just two hydrogen atoms attached to a single oxygen atom. Indeed, very few molecules are smaller or lighter. Liquid water, however, is the most extraordinary material contradicting its apparently simple molecular constituent. ⇒

Although we drink it, wash, fish, swim in it, and cook with it (although probably not all simultaneously), we nearly always overlook the special relationship it has with our lives. Droughts cause famines, and floods cause death and disease. It makes up over about half of us and, without it, we die within a few days. Liquid water has importance as a solvent, a solute, a reactant, a catalyst, and a biomolecule, structuring proteins, nucleic acid, and cells and controlling our consciousness. H 2 O is the second most common molecule in the Universe (behind hydrogen, H 2 ), the most abundant solid material and fundamental to star formation. As a hydride, it differs from other gaseous hydrides in not being poisonous or smelly. There is a hundred times as many water molecules in our bodies than the sum of all the other molecules put together, with billions of water molecules per DNA molecule. Life cannot evolve or continue without liquid water, which is why there is so much excitement about finding it on Mars and other planets and moons. Unsurprisingly, water plays a central role in many of the World's Religions. This website discusses many aspects of water science. ⇒

Water is the most studied material on Earth. However, it is remarkable to find that the science behind its behavior and function are so poorly understood (or even ignored), not only by people in general but also by scientists working with it every day. It can be extremely slippery and extremely sticky simultaneously, and this 'stick/slip' behavior is how we recognize the feel of water. The small size of its molecule belies the complexity of its actions and its unique capabilities. Many attempts to model water as a simple substance have failed and still are failing. Liquid water's unique properties and chameleonic nature seem to fit ideally into the requirements for life, as can no other molecule. ⇒

Many explanations of the complex behavior of liquid water have been published, with several stirring up considerable controversy. In this website, I have attempted to present and explain these ideas in a self-consistent and balanced manner, which I hope will encourage its understanding and further work. ⇒

Svalbard; gaseous, liquid and solid water

Water Site Map

Text, HTML markup, and images on this website are copyright. The maintenance of this site has no connection with commercial organizations or their products. Please do not copy from this site, or the occasional pirated copies of this site, without proper acknowledgment back to this site. Pirated copies can usually be distinguished by their outdated date stamps.

This site represents the author's views, mostly without prior peer review but supported by peer-reviewed papers. It attempts to give an unbiased, neutral point of view. Comments and criticisms are welcomed both by email and by using the Visitors book . I have always guaranteed to publish all submissions via the Visitor's book, only lightly editing their spelling and grammar and removing gratuitous advertisements, spam, or offensive material.

Calculations using the Restricted Hartree-Fock wave function (RHF) using the 6-31G** basis set make use of the Hyperchem Professional 8.0 package . To see all the animations, the viewer needs to be java enabled. Jmol is an open-source Java viewer for chemical structures in 3D. Some interactive pages use Chemical Objects for the Web (*.cow) or Jmol (*.pdb) files. The (*.cow) HyperChem Web Viewer (2 MB) is free from Hypercube, Inc. It seems that Internet Explorer 11 may be the easiest program to set up for javascript on Windows 10.

If you wish to consult previous versions of this website, there are 796 copies on the Wayback Internet Archive (as of 22 August 2020) archived between October 1 2000 and August 22 2020 ( https://web.archive.org/web/*/www.sbu.ac.uk/water/ and https://web.archive.org/web/*/www.lsbu.ac.uk/water/ ). Newer copies are to be found at . https://web.archive.org/web/20211109084920/https://water.lsbu.ac.uk/water/

As at 4 Jan 2022, this Water Structure and Science website has 54 MB, 315 webpages, 5,271 references, >30,000 links, >320,000 words, >1,800 images, and >120 3-D visualization and animation files. The mean readability is similar to that of the New York Times. The data in this Web site has stood the test of time, with chunks copied into Wikipedia and Chemical Reviews. There are over 2,800 scientific citations of this website, or its predecessor (lsbu.ac.uk/water), from Google Scholar. This website changed its address from www.lsbu.ac.uk/water on November 8, 2021 due to a situation beyond the author's control.

This page was established in 2000 and last updated by Martin Chaplin on 4 January, 2022

Creative Commons License

  • Search Menu

Sign in through your institution

  • Browse content in Arts and Humanities
  • Browse content in Archaeology
  • Anglo-Saxon and Medieval Archaeology
  • Archaeological Methodology and Techniques
  • Archaeology by Region
  • Archaeology of Religion
  • Archaeology of Trade and Exchange
  • Biblical Archaeology
  • Contemporary and Public Archaeology
  • Environmental Archaeology
  • Historical Archaeology
  • History and Theory of Archaeology
  • Industrial Archaeology
  • Landscape Archaeology
  • Mortuary Archaeology
  • Prehistoric Archaeology
  • Underwater Archaeology
  • Zooarchaeology
  • Browse content in Architecture
  • Architectural Structure and Design
  • History of Architecture
  • Residential and Domestic Buildings
  • Theory of Architecture
  • Browse content in Art
  • Art Subjects and Themes
  • History of Art
  • Industrial and Commercial Art
  • Theory of Art
  • Biographical Studies
  • Byzantine Studies
  • Browse content in Classical Studies
  • Classical History
  • Classical Philosophy
  • Classical Mythology
  • Classical Numismatics
  • Classical Literature
  • Classical Reception
  • Classical Art and Architecture
  • Classical Oratory and Rhetoric
  • Greek and Roman Epigraphy
  • Greek and Roman Law
  • Greek and Roman Archaeology
  • Greek and Roman Papyrology
  • Late Antiquity
  • Religion in the Ancient World
  • Social History
  • Digital Humanities
  • Browse content in History
  • Colonialism and Imperialism
  • Diplomatic History
  • Environmental History
  • Genealogy, Heraldry, Names, and Honours
  • Genocide and Ethnic Cleansing
  • Historical Geography
  • History by Period
  • History of Agriculture
  • History of Education
  • History of Emotions
  • History of Gender and Sexuality
  • Industrial History
  • Intellectual History
  • International History
  • Labour History
  • Legal and Constitutional History
  • Local and Family History
  • Maritime History
  • Military History
  • National Liberation and Post-Colonialism
  • Oral History
  • Political History
  • Public History
  • Regional and National History
  • Revolutions and Rebellions
  • Slavery and Abolition of Slavery
  • Social and Cultural History
  • Theory, Methods, and Historiography
  • Urban History
  • World History
  • Browse content in Language Teaching and Learning
  • Language Learning (Specific Skills)
  • Language Teaching Theory and Methods
  • Browse content in Linguistics
  • Applied Linguistics
  • Cognitive Linguistics
  • Computational Linguistics
  • Forensic Linguistics
  • Grammar, Syntax and Morphology
  • Historical and Diachronic Linguistics
  • History of English
  • Language Acquisition
  • Language Variation
  • Language Families
  • Language Evolution
  • Language Reference
  • Lexicography
  • Linguistic Theories
  • Linguistic Typology
  • Linguistic Anthropology
  • Phonetics and Phonology
  • Psycholinguistics
  • Sociolinguistics
  • Translation and Interpretation
  • Writing Systems
  • Browse content in Literature
  • Bibliography
  • Children's Literature Studies
  • Literary Studies (Asian)
  • Literary Studies (European)
  • Literary Studies (Eco-criticism)
  • Literary Studies (Modernism)
  • Literary Studies (Romanticism)
  • Literary Studies (American)
  • Literary Studies - World
  • Literary Studies (1500 to 1800)
  • Literary Studies (19th Century)
  • Literary Studies (20th Century onwards)
  • Literary Studies (African American Literature)
  • Literary Studies (British and Irish)
  • Literary Studies (Early and Medieval)
  • Literary Studies (Fiction, Novelists, and Prose Writers)
  • Literary Studies (Gender Studies)
  • Literary Studies (Graphic Novels)
  • Literary Studies (History of the Book)
  • Literary Studies (Plays and Playwrights)
  • Literary Studies (Poetry and Poets)
  • Literary Studies (Postcolonial Literature)
  • Literary Studies (Queer Studies)
  • Literary Studies (Science Fiction)
  • Literary Studies (Travel Literature)
  • Literary Studies (War Literature)
  • Literary Studies (Women's Writing)
  • Literary Theory and Cultural Studies
  • Mythology and Folklore
  • Shakespeare Studies and Criticism
  • Browse content in Media Studies
  • Browse content in Music
  • Applied Music
  • Dance and Music
  • Ethics in Music
  • Ethnomusicology
  • Gender and Sexuality in Music
  • Medicine and Music
  • Music Cultures
  • Music and Religion
  • Music and Culture
  • Music and Media
  • Music Education and Pedagogy
  • Music Theory and Analysis
  • Musical Scores, Lyrics, and Libretti
  • Musical Structures, Styles, and Techniques
  • Musicology and Music History
  • Performance Practice and Studies
  • Race and Ethnicity in Music
  • Sound Studies
  • Browse content in Performing Arts
  • Browse content in Philosophy
  • Aesthetics and Philosophy of Art
  • Epistemology
  • Feminist Philosophy
  • History of Western Philosophy
  • Metaphysics
  • Moral Philosophy
  • Non-Western Philosophy
  • Philosophy of Science
  • Philosophy of Action
  • Philosophy of Law
  • Philosophy of Religion
  • Philosophy of Language
  • Philosophy of Mind
  • Philosophy of Perception
  • Philosophy of Mathematics and Logic
  • Practical Ethics
  • Social and Political Philosophy
  • Browse content in Religion
  • Biblical Studies
  • Christianity
  • East Asian Religions
  • History of Religion
  • Judaism and Jewish Studies
  • Qumran Studies
  • Religion and Education
  • Religion and Health
  • Religion and Politics
  • Religion and Science
  • Religion and Law
  • Religion and Art, Literature, and Music
  • Religious Studies
  • Browse content in Society and Culture
  • Cookery, Food, and Drink
  • Cultural Studies
  • Customs and Traditions
  • Ethical Issues and Debates
  • Hobbies, Games, Arts and Crafts
  • Natural world, Country Life, and Pets
  • Popular Beliefs and Controversial Knowledge
  • Sports and Outdoor Recreation
  • Technology and Society
  • Travel and Holiday
  • Visual Culture
  • Browse content in Law
  • Arbitration
  • Browse content in Company and Commercial Law
  • Commercial Law
  • Company Law
  • Browse content in Comparative Law
  • Systems of Law
  • Competition Law
  • Browse content in Constitutional and Administrative Law
  • Government Powers
  • Judicial Review
  • Local Government Law
  • Military and Defence Law
  • Parliamentary and Legislative Practice
  • Construction Law
  • Contract Law
  • Browse content in Criminal Law
  • Criminal Procedure
  • Criminal Evidence Law
  • Sentencing and Punishment
  • Employment and Labour Law
  • Environment and Energy Law
  • Browse content in Financial Law
  • Banking Law
  • Insolvency Law
  • History of Law
  • Human Rights and Immigration
  • Intellectual Property Law
  • Browse content in International Law
  • Private International Law and Conflict of Laws
  • Public International Law
  • IT and Communications Law
  • Jurisprudence and Philosophy of Law
  • Law and Politics
  • Law and Society
  • Browse content in Legal System and Practice
  • Courts and Procedure
  • Legal Skills and Practice
  • Legal System - Costs and Funding
  • Primary Sources of Law
  • Regulation of Legal Profession
  • Medical and Healthcare Law
  • Browse content in Policing
  • Criminal Investigation and Detection
  • Police and Security Services
  • Police Procedure and Law
  • Police Regional Planning
  • Browse content in Property Law
  • Personal Property Law
  • Restitution
  • Study and Revision
  • Terrorism and National Security Law
  • Browse content in Trusts Law
  • Wills and Probate or Succession
  • Browse content in Medicine and Health
  • Browse content in Allied Health Professions
  • Arts Therapies
  • Clinical Science
  • Dietetics and Nutrition
  • Occupational Therapy
  • Operating Department Practice
  • Physiotherapy
  • Radiography
  • Speech and Language Therapy
  • Browse content in Anaesthetics
  • General Anaesthesia
  • Browse content in Clinical Medicine
  • Acute Medicine
  • Cardiovascular Medicine
  • Clinical Genetics
  • Clinical Pharmacology and Therapeutics
  • Dermatology
  • Endocrinology and Diabetes
  • Gastroenterology
  • Genito-urinary Medicine
  • Geriatric Medicine
  • Infectious Diseases
  • Medical Oncology
  • Medical Toxicology
  • Pain Medicine
  • Palliative Medicine
  • Rehabilitation Medicine
  • Respiratory Medicine and Pulmonology
  • Rheumatology
  • Sleep Medicine
  • Sports and Exercise Medicine
  • Clinical Neuroscience
  • Community Medical Services
  • Critical Care
  • Emergency Medicine
  • Forensic Medicine
  • Haematology
  • History of Medicine
  • Browse content in Medical Dentistry
  • Oral and Maxillofacial Surgery
  • Paediatric Dentistry
  • Restorative Dentistry and Orthodontics
  • Surgical Dentistry
  • Medical Ethics
  • Browse content in Medical Skills
  • Clinical Skills
  • Communication Skills
  • Nursing Skills
  • Surgical Skills
  • Medical Statistics and Methodology
  • Browse content in Neurology
  • Clinical Neurophysiology
  • Neuropathology
  • Nursing Studies
  • Browse content in Obstetrics and Gynaecology
  • Gynaecology
  • Occupational Medicine
  • Ophthalmology
  • Otolaryngology (ENT)
  • Browse content in Paediatrics
  • Neonatology
  • Browse content in Pathology
  • Chemical Pathology
  • Clinical Cytogenetics and Molecular Genetics
  • Histopathology
  • Medical Microbiology and Virology
  • Patient Education and Information
  • Browse content in Pharmacology
  • Psychopharmacology
  • Browse content in Popular Health
  • Caring for Others
  • Complementary and Alternative Medicine
  • Self-help and Personal Development
  • Browse content in Preclinical Medicine
  • Cell Biology
  • Molecular Biology and Genetics
  • Reproduction, Growth and Development
  • Primary Care
  • Professional Development in Medicine
  • Browse content in Psychiatry
  • Addiction Medicine
  • Child and Adolescent Psychiatry
  • Forensic Psychiatry
  • Learning Disabilities
  • Old Age Psychiatry
  • Psychotherapy
  • Browse content in Public Health and Epidemiology
  • Epidemiology
  • Public Health
  • Browse content in Radiology
  • Clinical Radiology
  • Interventional Radiology
  • Nuclear Medicine
  • Radiation Oncology
  • Reproductive Medicine
  • Browse content in Surgery
  • Cardiothoracic Surgery
  • Gastro-intestinal and Colorectal Surgery
  • General Surgery
  • Neurosurgery
  • Paediatric Surgery
  • Peri-operative Care
  • Plastic and Reconstructive Surgery
  • Surgical Oncology
  • Transplant Surgery
  • Trauma and Orthopaedic Surgery
  • Vascular Surgery
  • Browse content in Science and Mathematics
  • Browse content in Biological Sciences
  • Aquatic Biology
  • Biochemistry
  • Bioinformatics and Computational Biology
  • Developmental Biology
  • Ecology and Conservation
  • Evolutionary Biology
  • Genetics and Genomics
  • Microbiology
  • Molecular and Cell Biology
  • Natural History
  • Plant Sciences and Forestry
  • Research Methods in Life Sciences
  • Structural Biology
  • Systems Biology
  • Zoology and Animal Sciences
  • Browse content in Chemistry
  • Analytical Chemistry
  • Computational Chemistry
  • Crystallography
  • Environmental Chemistry
  • Industrial Chemistry
  • Inorganic Chemistry
  • Materials Chemistry
  • Medicinal Chemistry
  • Mineralogy and Gems
  • Organic Chemistry
  • Physical Chemistry
  • Polymer Chemistry
  • Study and Communication Skills in Chemistry
  • Theoretical Chemistry
  • Browse content in Computer Science
  • Artificial Intelligence
  • Computer Architecture and Logic Design
  • Game Studies
  • Human-Computer Interaction
  • Mathematical Theory of Computation
  • Programming Languages
  • Software Engineering
  • Systems Analysis and Design
  • Virtual Reality
  • Browse content in Computing
  • Business Applications
  • Computer Security
  • Computer Games
  • Computer Networking and Communications
  • Digital Lifestyle
  • Graphical and Digital Media Applications
  • Operating Systems
  • Browse content in Earth Sciences and Geography
  • Atmospheric Sciences
  • Environmental Geography
  • Geology and the Lithosphere
  • Maps and Map-making
  • Meteorology and Climatology
  • Oceanography and Hydrology
  • Palaeontology
  • Physical Geography and Topography
  • Regional Geography
  • Soil Science
  • Urban Geography
  • Browse content in Engineering and Technology
  • Agriculture and Farming
  • Biological Engineering
  • Civil Engineering, Surveying, and Building
  • Electronics and Communications Engineering
  • Energy Technology
  • Engineering (General)
  • Environmental Science, Engineering, and Technology
  • History of Engineering and Technology
  • Mechanical Engineering and Materials
  • Technology of Industrial Chemistry
  • Transport Technology and Trades
  • Browse content in Environmental Science
  • Applied Ecology (Environmental Science)
  • Conservation of the Environment (Environmental Science)
  • Environmental Sustainability
  • Environmentalist Thought and Ideology (Environmental Science)
  • Management of Land and Natural Resources (Environmental Science)
  • Natural Disasters (Environmental Science)
  • Nuclear Issues (Environmental Science)
  • Pollution and Threats to the Environment (Environmental Science)
  • Social Impact of Environmental Issues (Environmental Science)
  • History of Science and Technology
  • Browse content in Materials Science
  • Ceramics and Glasses
  • Composite Materials
  • Metals, Alloying, and Corrosion
  • Nanotechnology
  • Browse content in Mathematics
  • Applied Mathematics
  • Biomathematics and Statistics
  • History of Mathematics
  • Mathematical Education
  • Mathematical Finance
  • Mathematical Analysis
  • Numerical and Computational Mathematics
  • Probability and Statistics
  • Pure Mathematics
  • Browse content in Neuroscience
  • Cognition and Behavioural Neuroscience
  • Development of the Nervous System
  • Disorders of the Nervous System
  • History of Neuroscience
  • Invertebrate Neurobiology
  • Molecular and Cellular Systems
  • Neuroendocrinology and Autonomic Nervous System
  • Neuroscientific Techniques
  • Sensory and Motor Systems
  • Browse content in Physics
  • Astronomy and Astrophysics
  • Atomic, Molecular, and Optical Physics
  • Biological and Medical Physics
  • Classical Mechanics
  • Computational Physics
  • Condensed Matter Physics
  • Electromagnetism, Optics, and Acoustics
  • History of Physics
  • Mathematical and Statistical Physics
  • Measurement Science
  • Nuclear Physics
  • Particles and Fields
  • Plasma Physics
  • Quantum Physics
  • Relativity and Gravitation
  • Semiconductor and Mesoscopic Physics
  • Browse content in Psychology
  • Affective Sciences
  • Clinical Psychology
  • Cognitive Neuroscience
  • Cognitive Psychology
  • Criminal and Forensic Psychology
  • Developmental Psychology
  • Educational Psychology
  • Evolutionary Psychology
  • Health Psychology
  • History and Systems in Psychology
  • Music Psychology
  • Neuropsychology
  • Organizational Psychology
  • Psychological Assessment and Testing
  • Psychology of Human-Technology Interaction
  • Psychology Professional Development and Training
  • Research Methods in Psychology
  • Social Psychology
  • Browse content in Social Sciences
  • Browse content in Anthropology
  • Anthropology of Religion
  • Human Evolution
  • Medical Anthropology
  • Physical Anthropology
  • Regional Anthropology
  • Social and Cultural Anthropology
  • Theory and Practice of Anthropology
  • Browse content in Business and Management
  • Business Strategy
  • Business History
  • Business Ethics
  • Business and Government
  • Business and Technology
  • Business and the Environment
  • Comparative Management
  • Corporate Governance
  • Corporate Social Responsibility
  • Entrepreneurship
  • Health Management
  • Human Resource Management
  • Industrial and Employment Relations
  • Industry Studies
  • Information and Communication Technologies
  • International Business
  • Knowledge Management
  • Management and Management Techniques
  • Operations Management
  • Organizational Theory and Behaviour
  • Pensions and Pension Management
  • Public and Nonprofit Management
  • Social Issues in Business and Management
  • Strategic Management
  • Supply Chain Management
  • Browse content in Criminology and Criminal Justice
  • Criminal Justice
  • Criminology
  • Forms of Crime
  • International and Comparative Criminology
  • Youth Violence and Juvenile Justice
  • Development Studies
  • Browse content in Economics
  • Agricultural, Environmental, and Natural Resource Economics
  • Asian Economics
  • Behavioural Finance
  • Behavioural Economics and Neuroeconomics
  • Econometrics and Mathematical Economics
  • Economic Systems
  • Economic Methodology
  • Economic History
  • Economic Development and Growth
  • Financial Markets
  • Financial Institutions and Services
  • General Economics and Teaching
  • Health, Education, and Welfare
  • History of Economic Thought
  • International Economics
  • Labour and Demographic Economics
  • Law and Economics
  • Macroeconomics and Monetary Economics
  • Microeconomics
  • Public Economics
  • Urban, Rural, and Regional Economics
  • Welfare Economics
  • Browse content in Education
  • Adult Education and Continuous Learning
  • Care and Counselling of Students
  • Early Childhood and Elementary Education
  • Educational Equipment and Technology
  • Educational Strategies and Policy
  • Higher and Further Education
  • Organization and Management of Education
  • Philosophy and Theory of Education
  • Schools Studies
  • Secondary Education
  • Teaching of a Specific Subject
  • Teaching of Specific Groups and Special Educational Needs
  • Teaching Skills and Techniques
  • Browse content in Environment
  • Applied Ecology (Social Science)
  • Climate Change
  • Conservation of the Environment (Social Science)
  • Environmentalist Thought and Ideology (Social Science)
  • Management of Land and Natural Resources (Social Science)
  • Natural Disasters (Environment)
  • Pollution and Threats to the Environment (Social Science)
  • Social Impact of Environmental Issues (Social Science)
  • Sustainability
  • Browse content in Human Geography
  • Cultural Geography
  • Economic Geography
  • Political Geography
  • Browse content in Interdisciplinary Studies
  • Communication Studies
  • Museums, Libraries, and Information Sciences
  • Browse content in Politics
  • African Politics
  • Asian Politics
  • Chinese Politics
  • Comparative Politics
  • Conflict Politics
  • Elections and Electoral Studies
  • Environmental Politics
  • Ethnic Politics
  • European Union
  • Foreign Policy
  • Gender and Politics
  • Human Rights and Politics
  • Indian Politics
  • International Relations
  • International Organization (Politics)
  • Irish Politics
  • Latin American Politics
  • Middle Eastern Politics
  • Political Methodology
  • Political Communication
  • Political Philosophy
  • Political Sociology
  • Political Theory
  • Political Behaviour
  • Political Economy
  • Political Institutions
  • Politics and Law
  • Politics of Development
  • Public Administration
  • Public Policy
  • Qualitative Political Methodology
  • Quantitative Political Methodology
  • Regional Political Studies
  • Russian Politics
  • Security Studies
  • State and Local Government
  • UK Politics
  • US Politics
  • Browse content in Regional and Area Studies
  • African Studies
  • Asian Studies
  • East Asian Studies
  • Japanese Studies
  • Latin American Studies
  • Middle Eastern Studies
  • Native American Studies
  • Scottish Studies
  • Browse content in Research and Information
  • Research Methods
  • Browse content in Social Work
  • Addictions and Substance Misuse
  • Adoption and Fostering
  • Care of the Elderly
  • Child and Adolescent Social Work
  • Couple and Family Social Work
  • Direct Practice and Clinical Social Work
  • Emergency Services
  • Human Behaviour and the Social Environment
  • International and Global Issues in Social Work
  • Mental and Behavioural Health
  • Social Justice and Human Rights
  • Social Policy and Advocacy
  • Social Work and Crime and Justice
  • Social Work Macro Practice
  • Social Work Practice Settings
  • Social Work Research and Evidence-based Practice
  • Welfare and Benefit Systems
  • Browse content in Sociology
  • Childhood Studies
  • Community Development
  • Comparative and Historical Sociology
  • Disability Studies
  • Economic Sociology
  • Gender and Sexuality
  • Gerontology and Ageing
  • Health, Illness, and Medicine
  • Marriage and the Family
  • Migration Studies
  • Occupations, Professions, and Work
  • Organizations
  • Population and Demography
  • Race and Ethnicity
  • Social Theory
  • Social Movements and Social Change
  • Social Research and Statistics
  • Social Stratification, Inequality, and Mobility
  • Sociology of Religion
  • Sociology of Education
  • Sport and Leisure
  • Urban and Rural Studies
  • Browse content in Warfare and Defence
  • Defence Strategy, Planning, and Research
  • Land Forces and Warfare
  • Military Administration
  • Military Life and Institutions
  • Naval Forces and Warfare
  • Other Warfare and Defence Issues
  • Peace Studies and Conflict Resolution
  • Weapons and Equipment

The Structure and Properties of Water

  • < Previous
  • Next chapter >

The Structure and Properties of Water

1. The Water Molecule

  • Published: October 2005
  • Cite Icon Cite
  • Permissions Icon Permissions

This chapter presents descriptions of an isolated water molecule based on experiment and theory. The description based on experiment involves measurements made on water vapour at sufficiently low pressures or high temperatures to ensure that interactions between molecules are largely absent. The description based on theory provides details such as the shape of the electronic charge cloud of water, and an indication of which parts of the charge contribute most heavily to the total polarity of the molecule. The separation of these interdependent descriptions is artificial, but it serves to emphasize which portion of our understanding of water is based on observation, and which is based on reasonably accurate models of the molecule.

Personal account

  • Sign in with email/username & password
  • Get email alerts
  • Save searches
  • Purchase content
  • Activate your purchase/trial code
  • Add your ORCID iD

Institutional access

Sign in with a library card.

  • Sign in with username/password
  • Recommend to your librarian
  • Institutional account management
  • Get help with access

Access to content on Oxford Academic is often provided through institutional subscriptions and purchases. If you are a member of an institution with an active account, you may be able to access content in one of the following ways:

IP based access

Typically, access is provided across an institutional network to a range of IP addresses. This authentication occurs automatically, and it is not possible to sign out of an IP authenticated account.

Choose this option to get remote access when outside your institution. Shibboleth/Open Athens technology is used to provide single sign-on between your institution’s website and Oxford Academic.

  • Click Sign in through your institution.
  • Select your institution from the list provided, which will take you to your institution's website to sign in.
  • When on the institution site, please use the credentials provided by your institution. Do not use an Oxford Academic personal account.
  • Following successful sign in, you will be returned to Oxford Academic.

If your institution is not listed or you cannot sign in to your institution’s website, please contact your librarian or administrator.

Enter your library card number to sign in. If you cannot sign in, please contact your librarian.

Society Members

Society member access to a journal is achieved in one of the following ways:

Sign in through society site

Many societies offer single sign-on between the society website and Oxford Academic. If you see ‘Sign in through society site’ in the sign in pane within a journal:

  • Click Sign in through society site.
  • When on the society site, please use the credentials provided by that society. Do not use an Oxford Academic personal account.

If you do not have a society account or have forgotten your username or password, please contact your society.

Sign in using a personal account

Some societies use Oxford Academic personal accounts to provide access to their members. See below.

A personal account can be used to get email alerts, save searches, purchase content, and activate subscriptions.

Some societies use Oxford Academic personal accounts to provide access to their members.

Viewing your signed in accounts

Click the account icon in the top right to:

  • View your signed in personal account and access account management features.
  • View the institutional accounts that are providing access.

Signed in but can't access content

Oxford Academic is home to a wide variety of products. The institutional subscription may not cover the content that you are trying to access. If you believe you should have access to that content, please contact your librarian.

For librarians and administrators, your personal account also provides access to institutional account management. Here you will find options to view and activate subscriptions, manage institutional settings and access options, access usage statistics, and more.

Our books are available by subscription or purchase to libraries and institutions.

Month: Total Views:
October 2022 5
November 2022 10
December 2022 21
January 2023 26
February 2023 26
March 2023 6
April 2023 10
May 2023 4
June 2023 9
July 2023 6
September 2023 5
October 2023 16
November 2023 23
December 2023 14
January 2024 15
February 2024 12
March 2024 6
April 2024 2
May 2024 3
June 2024 5
July 2024 6
August 2024 5
  • About Oxford Academic
  • Publish journals with us
  • University press partners
  • What we publish
  • New features  
  • Open access
  • Rights and permissions
  • Accessibility
  • Advertising
  • Media enquiries
  • Oxford University Press
  • Oxford Languages
  • University of Oxford

Oxford University Press is a department of the University of Oxford. It furthers the University's objective of excellence in research, scholarship, and education by publishing worldwide

  • Copyright © 2024 Oxford University Press
  • Cookie settings
  • Cookie policy
  • Privacy policy
  • Legal notice

This Feature Is Available To Subscribers Only

Sign In or Create an Account

This PDF is available to Subscribers Only

For full access to this pdf, sign in to an existing account, or purchase an annual subscription.

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
  • Introduction & Top Questions
  • Liquid water
  • Structures of ice
  • Significance of the structure of liquid water
  • Water at high temperatures and pressures

Physical properties

Acid-base reactions, oxidation-reduction reactions.

water rapids, Niagara Falls, Canada

  • Where does water come from?
  • Why do cold water bottles and soft-drink bottles sweat?
  • When does water boil?
  • Why is water blue?
  • When is water the most dense?

Iceberg, Arctic (polar, environment, global warming)

Our editors will review what you’ve submitted and determine whether to revise the article.

  • Harvard T.H. Chan School of Public Health - The Nutrition Source - Water
  • Energy Education - Water
  • USGS - Water Science School - pH and Water
  • National Center for Biotechnology Information - PubChem - Water
  • Chemistry LibreTexts - All About Water
  • USGS Water Science School - Surface Tension and Water
  • water - Children's Encyclopedia (Ages 8-11)
  • water - Student Encyclopedia (Ages 11 and up)
  • Table Of Contents

Water has several important physical properties. Although these properties are familiar because of the omnipresence of water, most of the physical properties of water are quite atypical . Given the low molar mass of its constituent molecules, water has unusually large values of viscosity, surface tension , heat of vaporization , and entropy of vaporization, all of which can be ascribed to the extensive hydrogen bonding interactions present in liquid water. The open structure of ice that allows for maximum hydrogen bonding explains why solid water is less dense than liquid water—a highly unusual situation among common substances.

Selected physical properties of water
molar mass 18.0151 grams per mole
melting point 0.00 °C
boiling point 100.00 °C
maximum density (at 3.98 °C) 1.0000 grams per cubic centimetre
density (25 °C) 0.99701 grams per cubic centimetre
vapour pressure (25 °C) 23.75 torr
heat of fusion (0 °C) 6.010 kilojoules per mole
heat of vaporization (100 °C) 40.65 kilojoules per mole
heat of formation (25 °C) −285.85 kilojoules per mole
entropy of vaporization (25 °C) 118.8 joules per °C mole
viscosity 0.8903 centipoise
surface tension (25 °C) 71.97 dynes per centimeter

Chemical properties

Recent news.

Water undergoes various types of chemical reactions. One of the most important chemical properties of water is its ability to behave as both an acid (a proton donor) and a base (a proton acceptor), the characteristic property of amphoteric substances. This behaviour is most clearly seen in the autoionization of water: H 2 O(l) + H 2 O(l) ⇌ H 3 O + (aq) + OH − (aq), where the (l) represents the liquid state, the (aq) indicates that the species are dissolved in water, and the double arrows indicate that the reaction can occur in either direction and an equilibrium condition exists. At 25 °C (77 °F) the concentration of hydrated H + (i.e., H 3 O + , known as the hydronium ion) in water is 1.0 × 10 −7 M, where M represents moles per litre . Since one OH − ion is produced for each H 3 O + ion, the concentration of OH − at 25 °C is also 1.0 × 10 −7 M. In water at 25 °C the H 3 O + concentration and the OH − concentration must always be 1.0 × 10 −14 : [H + ][OH − ] = 1.0 × 10 −14 , where [H + ] represents the concentration of hydrated H + ions in moles per litre and [OH − ] represents the concentration of OH − ions in moles per litre.

When an acid (a substance that can produce H + ions) is dissolved in water, both the acid and the water contribute H + ions to the solution. This leads to a situation in which the H + concentration is greater than 1.0 × 10 −7 M. Since it must always be true that [H + ][OH − ] = 1.0 × 10 −14 at 25 °C, the [OH − ] must be lowered to some value below 1.0 × 10 −7 . The mechanism for reducing the concentration of OH − involves the reaction H + + OH − → H 2 O, which occurs to the extent needed to restore the product of [H + ] and [OH − ] to 1.0 × 10 −14 M. Thus, when an acid is added to water, the resulting solution contains more H + than OH − ; that is, [H + ] > [OH − ]. Such a solution (in which [H + ] > [OH − ]) is said to be acidic.

The most common method for specifying the acidity of a solution is its pH , which is defined in terms of the hydrogen ion concentration: pH = −log [H + ], where the symbol log stands for a base-10 logarithm . In pure water, in which [H + ] = 1.0 × 10 −7 M, the pH = 7.0. For an acidic solution, the pH is less than 7. When a base (a substance that behaves as a proton acceptor) is dissolved in water, the H + concentration is decreased so that [OH − ] > [H + ]. A basic solution is characterized by having a pH > 7. In summary, in aqueous solutions at 25 °C:

neutral solution [H ] = [OH ] pH = 7
acidic solution [H ] > [OH ] pH < 7
basic solution [OH ] > [H ] pH > 7

Lake Mead is seen in the distance behind a dead creosote bush in an area of dry, cracked earth that used to be underwater near where the Lake Mead Marina was once located on June 12, 2021 in the Lake Mead National Recreation Area, Nevada.

Noble metals, such as gold and silver , do not react with water at all.

Understanding the Structure and Function of Water at the Molecular Scale

  • First Online: 20 June 2023

Cite this chapter

water molecule essay

  • Sheng Meng 3 &
  • Enge Wang 4  

420 Accesses

If you had to choose “what is the most important substance in nature”, I’m afraid most people’s answer would be: water! This is probably one of the beliefs that have remained constant throughout the history of mankind. We have to ask: What is it about water and why is it so magical?

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

J.W. Döbereiner, Schweigg. J. 39 , 1 (1823)

Google Scholar  

J.J. Berzelius, Jber. Chem. 15 , 237 (1837)

U. Raviv, P. Laurat, J. Klein, Fluidity of water confined to sub nanometer films. Nature 413 , 51 (2001)

Article   ADS   Google Scholar  

D.D. Klug, Dense ice in detail. Nature 420 , 749 (2002)

O. Mishima, Y. Suzuki, Propagation of the polyamorphic transition of ice and the liquid-liquid critical point. Nature 419 , 599 (2002)

C.A. Tulk, C.J. Benmore, J. Urquidi, D.D. Klug, J. Neuefeind, B. Tomberli, P.A. Egelstaff, Structural studies of several distinct metastable forms of amorphous ice. Science 297 , 1320 (2002)

J. Li, D.K. Ross, Evidence for two kinds of hydrogen bond in ice. Nature 365 , 327 (1993)

W.L. Jorgensen, L. Salem, The Organic Chemists’ Book of Orbitals (Academic Press, New York, 1973)

T.R. Dyke, K.M. Mack, J.S. Muenter, The structure of water dimer from molecular beam electric resonance spectroscopy. J. Chem. Phys. 66 , 498 (1977)

N. Pugliano, R.J. Saykally, Measurement of quantum tunneling between chiral isomers of the cyclic water trimer. Science 257 , 193 (1992)

Article   Google Scholar  

J.D. Cruzan, L.B. Braly, K. Liu, M.G. Brown, J.G. Loeser, R.J. Saykally, Quantifying hydrogen bond cooperativity in water: VRT spectroscopy of the water tetramer. Science 271 , 59 (1996)

K. Liu, M.G. Brown, J.D. Cruzan, R.J. Saykally, Vibration-rotation tunneling spectra of the water pentamer: structure and dynamics. Science 271 , 62 (1996)

K. Liu, M.G. Brown, C. Carter, R.J. Saykally, J.K. Gregory, D.C. Clary, Characterization of a cage form of the water hexamer. Nature 381 , 501 (1996)

K. Liu, J.D. Cruzan, R.J. Saykally, Water clusters. Science 271 , 929 (1996)

J.K. Gregory, D.C. Clary, K. Liu, M.G. Brown, R.J. Saykally, The water dipole moment in water clusters. Science 275 , 814 (1997)

J.D. Bernal, R.H. Fowler, A theory of water and ionic solution, with particular reference to hydrogen and hydroxyl ions. J. Chem. Phys. 1 , 515 (1933)

L. Pauling, The structure and entropy of ice and of other crystals with some randomness of atomic arrangement. J. Am. Chem. Soc. 57 , 2680 (1935)

K. Modig, B.G. Pfrommer, B. Halle, Temperature-dependent hydrogen-bond geometry in liquid water. Phys. Rev. Lett. 90 , 075502 (2003)

M. Boero, K. Terakura, T. Ikeshoji, C.C. Liew, M. Parrinello, Hydrogen bonding and dipole moment of water at supercritical conditions: a first-principles molecular dynamics study. Phys. Rev. Lett. 85 , 3245 (2000)

E. Schwegler, G. Galli, F. Gygi, R.Q. Hood, Dissociation of water under pressure. Phys. Rev. Lett. 87 , 265501 (2001)

M.F. Kropman, H.J. Bakker, Dynamics of water molecules in aqueous solvation shells. Science 291 , 2118 (2001)

D. Marx, M.E. Tuckerman, J. Hutter, M. Parrinello, The nature of the hydrated excess proton in water. Nature 397 , 601 (1999)

M.E. Tuckerman, D. Marx, M. Parrinello, The nature and transport mechanism of hydrated hydroxide ions in aqueous solution. Nature 417 , 925 (2002)

J. Alper, Water splitting goes Au naturel. Science 299 , 1686 (2003)

F.L. Gervasio, P. Carloni, M. Parrinello, Electronic structure of wet DNA. Phys. Rev. Lett. 89 , 108102 (2002)

K. Laasonen, M. Sprik, M. Parrinello, “Ab initio” liquid water. J. Chem. Phys. 99 , 9080 (1993)

M.E. Tuckerman, D. Marx, M.L. Klein, M. Parrinello, On the quantum nature of the shared proton in hydrogen bonds. Science 275 , 817 (1997)

M. Benoit, D. Marx, M. Parrinello, Quantum effects on phase transitions in high-pressure ice. Comput. Mater. Sci. 10 , 88 (1998)

L.E. Firment, G.A. Somorjai, Low-energy electron diffraction studies of molecular crystals: the surface structures of vapor-grown ice and naphthalene. J. Chem. Phys. 63 , 1037 (1975)

L.E. Firment, G.A. Somorjai, Low energy electron diffraction studies of the surfaces of molecular crystals (ice, ammonia, naphthalene, benzene). Surf. Sci. 84 , 275 (1979)

B.A. Sexton, Vibrational spectra of water chemisorbed on Platinum (111). Surf. Sci. 94 , 435 (1980)

G.B. Fisher, J.L. Gland, The interaction of water with the Pt(111) surface. Surf. Sci. 94 , 446 (1980)

D. Doering, T.E. Madey, The adsorption of water on clean and oxygen-dosed Ru(001). Surf. Sci. 123 , 305 (1982)

P.A. Thiel, R.A. DePaola, F.M. Hoffmann, The vibrational spectra of chemisorbed molecular clusters: H 2 O on Ru(001). J. Chem. Phys. 80 , 5326 (1984)

S. Andersson, C. Nyberg, C.G. Tengstål, Adsorption of water monomers on Cu(100) and Pd(100) at low temperatures. Chem. Phys. Lett. 104 , 305 (1984)

F.T. Wagner, T.E. Moylan, A comparison between water adsorbed on Rh(111) and Pt(111), with and without predosed oxygen. Surf. Sci. 191 , 121 (1987)

W. Ranke, Low temperature adsorption and condensation of O 2 , H 2 O, and NO on Pt(111), studied by core level and valence band photoemission. Surf. Sci. 209 , 57 (1989)

A.F. Carley, P.R. Davies, M.W. Roberts, K.K. Thomas, Hydroxylation of molecularly adsorbed water at Ag(111) and Cu(100) surfaces by dioxygen: photoelectron and vibrational spectroscopic studies. Surf. Sci. 238 , L467 (1990)

H. Ogasawara, J. Yoshinobu, M. Kawai, Water adsorption on Pt(111): from isolated molecule to three-dimensional cluster. Chem. Phys. Lett. 231 , 188 (1994)

G. Held, D. Menzel, The structure of the p(√3×√3)R30° bilayer of D 2 O on Ru(001). Surf. Sci. 316 , 92 (1994)

M. Morgenstern, T. Michely, G. Comsa, Anisotropy in the adsorption of H 2 O at low coordination sites on Pt(111). Phys. Rev. Lett. 77 , 703 (1996)

M. Morgenstern, J. Muller, T. Michely, G. Comsa, The ice bilayer on Pt(111): nucleation, structure and melting. Z. Phys. Chem. 198 , 43 (1997)

I. Villegas, M.J. Weaver, Infrared spectroscopy of model electrochemical interfaces in ultrahigh vacuum: evidence for coupled cation-anion hydration in the Pt(111)/K + Cl − system. J. Phys. Chem. 100 , 19502 (1996)

S. Smith, C. Huang, E.K.L. Wong, B.D. Kay, Desorption and crystallization kinetics in nanoscale thin films of amorphous water ice. Surf. Sci. 367 , L13 (1996)

P. Löfgren, P. Ahlström, D.V. Chakarov, J. Lausmaa, B. Kasemo, Substrate dependent sublimation kinetics of mesoscopic ice films. Surf. Sci. 367 , L19 (1996)

A. Glebov, A.P. Graham, A. Menzel, J.P. Toennies, Orientational ordering of two-dimensional ice on Pt(111). J. Chem. Phys. 106 , 9382 (1997)

G. Pirug, H.P. Bonzel, UHV simulation of the electrochemical double layer: adsorption of HClO 4 /H 2 O on Au(111). Surf. Sci. 405 , 87 (1998)

X. Su, L. Lianos, Y.R. Shen, G.A. Somorjai, Surface-induced ferroelectric ice on Pt(111). Phys. Rev. Lett. 80 , 1533 (1998)

M.S. Yeganeh, S.M. Dougal, H.S. Pink, Vibrational spectroscopy of water at liquid/solid interfaces: crossing the isoelectric point of a solid surface. Phys. Rev. Lett. 83 , 1179 (1999)

H. Ogasawara, J. Yoshinobu, M. Kawai, Clustering behavior of water (D 2 O) on Pt(111). J. Chem. Phys. 111 , 7003 (1999)

M. Nakamura, Y. Shingaya, M. Ito, The vibrational spectra of water cluster molecules on Pt(111) surface at 20 K. Chem. Phys. Lett. 309 , 123 (1999)

A.L. Glebov, A.P. Graham, A. Menzel, Vibrational spectroscopy of water molecules on Pt(111) at submonolayer coverages. Surf. Sci. 427 , 22 (1999)

M. Nakamura, M. Ito, Monomer and tetramer water clusters adsorbed on Ru(0001). Chem. Phys. Lett. 325 , 293 (2000)

K. Jacobi, K. Bedürftig, Y. Wang, G. Ertl, From monomers to ice ─ new vibrational characteristics of H 2 O adsorbed on Pt(111). Surf. Sci. 472 , 9 (2001)

S. Haq, J. Harnett, A. Hodgson, Growth of thin crystalline ice films on Pt(111). Surf. Sci. 505 , 171 (2002)

T. Mitsui, M.K. Rose, E. Fomin, D.F. Ogletree, M. Salmeron, Water diffusion and clustering on Pd(111). Science 297 , 1850 (2002)

K. Morgenstern, J. Nieminen, Intermolecular bond length of ice on Ag(111). Phys. Rev. Lett. 88 , 066102 (2002)

H. Ogasawara, B. Brena, D. Nordlund, M. Nyberg, A. Pelmenschikov, L.G.M. Pettersson, A. Nilsson, Structure and bonding of water on Pt(111). Phys. Rev. Lett. 89 , 276102 (2002)

D.N. Denzler, C. Hess, R. Dudek, S. Wagner, C. Frischkorn, M. Wolf, G. Ertl, Interfacial structure of water on Ru(0001) investigated by vibrational spectroscopy. Chem. Phys. Lett. 376 , 618 (2003)

S.K. Jo, J. Kiss, J.A. Polanco, J.M. White, Identification of second layer adsorbates: water and chloroethane on Pt(111). Surf. Sci. 253 , 233 (1991)

P.J. Feibelman, Partial dissociation of water on Ru(0001). Science 295 , 99 (2002)

D.J. Elliot, Integrated Circuit Fabrication Technology (McGraw-Hill, New York, 1982)

W. Ranke, D. Schmeisser, Y.R. Xing, Orientation dependence of H 2 O adsorption on a cylindrical Si single crystal. Surf. Sci. 152–153 , 1103 (1985)

Y.J. Chabal, S.B. Christman, Evidence of dissociation of water on the Si(100)2×1 surface. Phys. Rev. B 29 , 6974 (1984)

R.A. Rosenberg, P.J. Love, V. Rehn, I. Owen, G. Thornton, The bonding of hydrogen on water-dosed Si(111). J. Vacuum Sci. Technol. A 4 , 1451 (1986)

F. Meyer, M.J. Sparnaay, The chemical and physical properties of clean Germanium and Silicon surfaces, in Surface physics of phosphors and semiconductors , ed. by C.G. Scott, C.E. Reed (Academic Press, London, 1975), chap. 6, pp. 321

R.H. Williams, I.T. McGovern, Adsorption on semiconductors, in the chemical physics of solid surfaces and heterogeneous catalysis, vol. 3, ed. by D.A. King, D.P. Woodruff (Elsevier, Amsterdam, 1984), chap. 6, pp. 267

W. Langel, M. Parrinello, Hydrolysis at stepped MgO surfaces. Phys. Rev. Lett. 73 , 504 (1994)

L. Giordano, J. Goniakowski, J. Suzanne, Partial dissociation of water molecules in the (3×2) water monolayer deposited on the MgO(100) surface. Phys. Rev. Lett. 81 , 1271 (1998)

M. Odelius, Mixed molecular and dissociative water adsorption on MgO[100]. Phys. Rev. Lett. 82 , 3919 (1999)

K.C. Hass, W.F. Schneider, A. Curioni, W. Andreoni, The chemistry of water on alumina surfaces: reaction dynamics from first principles. Science 282 , 265 (1998)

P.J.D. Lindan, N.M. Harrison, Mixed dissociative and molecular adsorption of water on the rutile (110) surface. Phys. Rev. Lett. 80 , 762 (1998)

A. Vittadini, A. Selloni, F.P. Rotzinger, M. Gratzel, Structure and energetics of water adsorbed at TiO 2 anatase (101) and (001) surfaces. Phys. Rev. Lett. 81 , 2954 (1998)

R. Schaub et al., Oxygen vacancies as active sites for water dissociation on rutile TiO 2 (110). Phys. Rev. Lett. 87 , 266104 (2001)

Y.S. Chu, T.E. Lister, W.G. Cullen, H. You, Z. Nagy, Commensurate water monolayer at the RuO 2 (110) water interface. Phys. Rev. Lett. 86 , 3364 (2001)

M. Odelius, M. Bernasconi, M. Parrinello, Two dimensional ice adsorbed on mica surface. Phys. Rev. Lett. 78 , 2855 (1997)

L. Cheng, P. Fenter, K.L. Nagy, M.L. Schlegel, N.C. Sturchio, Molecular-scale density oscillations in water adjacent to a mica surface. Phys. Rev. Lett. 87 , 156103 (2001)

S. Fölsch, M. Henzler, Water adsorption on the NaCl surface. Surf. Sci. 247 , 269 (1991)

S. Fölsch, A. Stock, M. Henzler, Two-dimensional water condensation on the NaCl(100) surface. Surf. Sci. 264 , 65 (1992)

L.W. Bruch, A. Glebov, J.P. Toennies, H. Weiss, A helium atom scattering study of water adsorption on the NaCl(100) single crystal surface. J. Chem. Phys. 103 , 5109 (1995)

M. Foster, G.E. Ewing, An infrared spectroscopic study of water thin films on NaCl(100). Surf. Sci. 427 , 102 (1999)

A. Allouche, Water adsorption on NaCl(100): a quantum ab-initio cluster calculation. Surf. Sci. 406 , 279 (1998)

H. Meyer, P. Entel, J. Hafner, Physisorption of water on salt surfaces. Surf. Sci. 488 , 177 (2001)

M.F. Reedijk, J. Arsic, F.F.A. Hollander, S.A. de Vries, E. Vlieg, Liquid order at the interface of KDP crystals with water: evidence for icelike layers. Phys. Rev. Lett. 90 , 066103 (2003)

D. Chakarov, B. Kasemo, Photoinduced crystallization of amorphous ice films on graphite. Phys. Rev. Lett. 81 , 5181 (1998)

T. Werder, J.H. Walther, R.L. Jaffe, T. Halicioglu, P. Koumoutsakos, On the water–carbon interaction for use in molecular dynamics simulations of graphite. J. Phys. Chem. B 107 , 1345 (2003)

L.F. Scatena, M.G. Brown, G.L. Richmond, Water at hydrophobic surfaces: weak hydrogen bonding and strong orientation effects. Science 292 , 908 (2001)

T.R. Jensen, M.O. Jensen, N. Reitzel, K. Balashev, G.H. Peters, K. Kjaer, T. Bjornholm, Water in contact with extended hydrophobic surfaces: direct evidence of weak dewetting. Phys. Rev. Lett. 90 , 086101 (2003)

L. Yang, H.W. Huang, Observation of a memberane fusion intermediate structure. Science 297 , 1877 (2002)

L. Guidoni, V. Torre, P. Carloni, Potassium and sodium binding to the outer mouth of the K + channel. Biochemistry 38 , 8599 (1999)

K. Murata et al., Structural determinants of water permeation through aquaporin-1. Nature 407 , 599 (2000)

P.A. Thiel, T.E. Madey, The interation of water with solid surfaces: fundamental aspects. Surf. Sci. Rep. 7 , 211 (1987)

M.A. Henderson, The interation of water with solid surfaces: fundamental aspects revisited. Surf. Sci. Rep. 46 , 1 (2002)

A. Hodgson, S. Haq, Water adsorption and the wetting of metal surfaces. Surf. Sci. Rep. 64 , 381 (2009)

A. Verdaguer, G.M. Sacha, H. Bluhm, M. Salmeron, Molecular structure of water at interfaces: wetting at the nanometer scale. Chem. Rev. 106 , 1478 (2006)

N.E. Levinger, Water in confinement. Science 298 , 1722 (2002)

P. Ball, How to keep dry in water. Nature 423 , 25 (2003)

J. Teixeira, J.M. Zanotti, M.C. BellissentFunel, S.H. Chen, Water in confined geometries. Phys. B 234 , 370 (1997)

R. Bergman, J. Swenson, Dynamics of supercooled water in confined geometry. Nature 403 , 283 (2000)

K. Koga, G.T. Gao, H. Tanaka, X.C. Zeng, Formation of ordered ice nanotubes inside carbon nanotubes. Nature 412 , 802 (2001)

G. Hummer, J.C. Rasaiah, J.P. Noworyta, Water conduction through the hydrophobic channel of a carbon nanotube. Nature 414 , 188 (2001)

M.S.P. Sansom, P.C. Biggin, Water at the nanoscale. Nature 414 , 156 (2001)

C. Dellago, M.M. Naor, G. Hummer, Proton transport through water-filled carbon nanotubes. Phys. Rev. Lett. 90 , 105902 (2003)

D.J. Mann, M.D. Halls, Water alignment and proton conduction inside carbon nanotubes. Phys. Rev. Lett. 90 , 195503 (2003)

R.P. Feynman, There is plenty of room at the bottom, in Annual Meeting of the American Physical Society . 29th Dec, 1959. http://www.zyvex.com/nanotech/feynman.html

Download references

Author information

Authors and affiliations.

Institute of Physics, Chinese Academy of Sciences, Beijing, China

Peking University, Beijing, China

You can also search for this author in PubMed   Google Scholar

Corresponding author

Correspondence to Sheng Meng .

Rights and permissions

Reprints and permissions

Copyright information

© 2023 Peking University Press

About this chapter

Meng, S., Wang, E. (2023). Understanding the Structure and Function of Water at the Molecular Scale. In: Water. Springer, Singapore. https://doi.org/10.1007/978-981-99-1541-5_2

Download citation

DOI : https://doi.org/10.1007/978-981-99-1541-5_2

Published : 20 June 2023

Publisher Name : Springer, Singapore

Print ISBN : 978-981-99-1540-8

Online ISBN : 978-981-99-1541-5

eBook Packages : Physics and Astronomy Physics and Astronomy (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

Logo for BCcampus Open Publishing

Want to create or adapt books like this? Learn more about how Pressbooks supports open publishing practices.

Chapter 2: Introduction to the Chemistry of Life

Learning Objectives

By the end of this section, you will be able to:

  • Describe the properties of water that are critical to maintaining life

Watch a video about why we need oxygen and how it causes problems for living things.

Do you ever wonder why scientists spend time looking for water on other planets? It is because water is essential to life; even minute traces of it on another planet can indicate that life could or did exist on that planet. Water is one of the more abundant molecules in living cells and the one most critical to life as we know it. Approximately 60–70 percent of your body is made up of water. Without it, life simply would not exist.

Water Is Polar

The hydrogen and oxygen atoms within water molecules form polar covalent bonds. The shared electrons spend more time associated with the oxygen atom than they do with hydrogen atoms. There is no overall charge to a water molecule, but there is a slight positive charge on each hydrogen atom and a slight negative charge on the oxygen atom. Because of these charges, the slightly positive hydrogen atoms repel each other and form the unique shape. Each water molecule attracts other water molecules because of the positive and negative charges in the different parts of the molecule. Water also attracts other polar molecules (such as sugars), forming hydrogen bonds. When a substance readily forms hydrogen bonds with water, it can dissolve in water and is referred to as hydrophilic (“water-loving”). Hydrogen bonds are not readily formed with nonpolar substances like oils and fats . These nonpolar compounds are hydrophobic (“water-fearing”) and will not dissolve in water.

Picture of oil in water.

Water Stabilizes Temperature

The hydrogen bonds in water allow it to absorb and release heat energy more slowly than many other substances. Temperature is a measure of the motion (kinetic energy) of molecules. As the motion increases, energy is higher and thus temperature is higher. Water absorbs a great deal of energy before its temperature rises. Increased energy disrupts the hydrogen bonds between water molecules. Because these bonds can be created and disrupted rapidly, water absorbs an increase in energy and temperature changes only minimally. This means that water moderates temperature changes within organisms and in their environments. As energy input continues, the balance between hydrogen-bond formation and destruction swings toward the destruction side. More bonds are broken than are formed. This process results in the release of individual water molecules at the surface of the liquid (such as a body of water, the leaves of a plant, or the skin of an organism) in a process called evaporation . Evaporation of sweat, which is 90 percent water, allows for cooling of an organism, because breaking hydrogen bonds requires an input of energy and takes heat away from the body.

Conversely, as molecular motion decreases and temperatures drop, less energy is present to break the hydrogen bonds between water molecules. These bonds remain intact and begin to form a rigid, lattice-like structure (e.g., ice) (Figure 2.8 a ). When frozen, ice is less dense than liquid water (the molecules are farther apart). This means that ice floats on the surface of a body of water (Figure 2.8 b ). In lakes, ponds, and oceans, ice will form on the surface of the water, creating an insulating barrier to protect the animal and plant life beneath from freezing in the water. If this did not happen, plants and animals living in water would freeze in a block of ice and could not move freely, making life in cold temperatures difficult or impossible.

Part A shows the lattice-like molecular structure of ice. Part B is a photo of ice on water.

Water Is an Excellent Solvent

Because water is polar, with slight positive and negative charges, ionic compounds and polar molecules can readily dissolve in it. Water is, therefore, what is referred to as a solvent —a substance capable of dissolving another substance. The charged particles will form hydrogen bonds with a surrounding layer of water molecules. This is referred to as a sphere of hydration and serves to keep the particles separated or dispersed in the water. In the case of table salt (NaCl) mixed in water, the sodium and chloride ions separate, or dissociate, in the water, and spheres of hydration are formed around the ions. A positively charged sodium ion is surrounded by the partially negative charges of oxygen atoms in water molecules. A negatively charged chloride ion is surrounded by the partially positive charges of hydrogen atoms in water molecules. These spheres of hydration are also referred to as hydration shells. The polarity of the water molecule makes it an effective solvent and is important in its many roles in living systems.

Illustration of spheres of hydration around sodium and chlorine ions.

Water Is Cohesive

Have you ever filled up a glass of water to the very top and then slowly added a few more drops? Before it overflows, the water actually forms a dome-like shape above the rim of the glass. This water can stay above the glass because of the property of cohesion . In cohesion, water molecules are attracted to each other (because of hydrogen bonding), keeping the molecules together at the liquid-air (gas) interface, although there is no more room in the glass. Cohesion gives rise to surface tension , the capacity of a substance to withstand rupture when placed under tension or stress. When you drop a small scrap of paper onto a droplet of water, the paper floats on top of the water droplet, although the object is denser (heavier) than the water. This occurs because of the surface tension that is created by the water molecules. Cohesion and surface tension keep the water molecules intact and the item floating on the top. It is even possible to “float” a steel needle on top of a glass of water if you place it gently, without breaking the surface tension.

Picture of a needle floating on top of water because of cohesion and surface tension.

These cohesive forces are also related to the water’s property of adhesion , or the attraction between water molecules and other molecules. This is observed when water “climbs” up a straw placed in a glass of water. You will notice that the water appears to be higher on the sides of the straw than in the middle. This is because the water molecules are attracted to the straw and therefore adhere to it.

Cohesive and adhesive forces are important for sustaining life. For example, because of these forces, water can flow up from the roots to the tops of plants to feed the plant.

Concept in Action

To learn more about water, visit the U.S. Geological Survey Water Science for Schools: All About Water! website.

Buffers, pH, Acids, and Bases

The pH of a solution is a measure of its acidity or alkalinity. You have probably used litmus paper , paper that has been treated with a natural water-soluble dye so it can be used as a pH indicator, to test how much acid or base (alkalinity) exists in a solution. You might have even used some to make sure the water in an outdoor swimming pool is properly treated. In both cases, this pH test measures the amount of hydrogen ions that exists in a given solution. High concentrations of hydrogen ions yield a low pH, whereas low levels of hydrogen ions result in a high pH. The overall concentration of hydrogen ions is inversely related to its pH and can be measured on the pH scale (Figure 2.11). Therefore, the more hydrogen ions present, the lower the pH; conversely, the fewer hydrogen ions, the higher the pH.

The pH scale ranges from 0 to 14. A change of one unit on the pH scale represents a change in the concentration of hydrogen ions by a factor of 10, a change in two units represents a change in the concentration of hydrogen ions by a factor of 100. Thus, small changes in pH represent large changes in the concentrations of hydrogen ions. Pure water is neutral. It is neither acidic nor basic, and has a pH of 7.0. Anything below 7.0 (ranging from 0.0 to 6.9) is acidic, and anything above 7.0 (from 7.1 to 14.0) is alkaline. The blood in your veins is slightly alkaline (pH = 7.4). The environment in your stomach is highly acidic (pH = 1 to 2). Orange juice is mildly acidic (pH = approximately 3.5), whereas baking soda is basic (pH = 9.0).

The pH scale with representative substances and their pHs.

Acids are substances that provide hydrogen ions (H + ) and lower pH, whereas bases provide hydroxide ions (OH – ) and raise pH. The stronger the acid, the more readily it donates H + . For example, hydrochloric acid and lemon juice are very acidic and readily give up H + when added to water. Conversely, bases are those substances that readily donate OH – . The OH – ions combine with H + to produce water, which raises a substance’s pH. Sodium hydroxide and many household cleaners are very alkaline and give up OH – rapidly when placed in water, thereby raising the pH.

Most cells in our bodies operate within a very narrow window of the pH scale, typically ranging only from 7.2 to 7.6. If the pH of the body is outside of this range, the respiratory system malfunctions, as do other organs in the body. Cells no longer function properly, and proteins will break down. Deviation outside of the pH range can induce coma or even cause death.

So how is it that we can ingest or inhale acidic or basic substances and not die? Buffers are the key. Buffers readily absorb excess H + or OH – , keeping the pH of the body carefully maintained in the aforementioned narrow range. Carbon dioxide is part of a prominent buffer system in the human body; it keeps the pH within the proper range. This buffer system involves carbonic acid (H 2 CO 3 ) and bicarbonate (HCO 3 – ) anion. If too much H + enters the body, bicarbonate will combine with the H + to create carbonic acid and limit the decrease in pH. Likewise, if too much OH – is introduced into the system, carbonic acid will rapidly dissociate into bicarbonate and H + ions. The H + ions can combine with the OH – ions, limiting the increase in pH. While carbonic acid is an important product in this reaction, its presence is fleeting because the carbonic acid is released from the body as carbon dioxide gas each time we breathe. Without this buffer system, the pH in our bodies would fluctuate too much and we would fail to survive.

Section Summary

Water has many properties that are critical to maintaining life. It is polar, allowing for the formation of hydrogen bonds, which allow ions and other polar molecules to dissolve in water. Therefore, water is an excellent solvent. The hydrogen bonds between water molecules give water the ability to hold heat better than many other substances. As the temperature rises, the hydrogen bonds between water continually break and reform, allowing for the overall temperature to remain stable, although increased energy is added to the system. Water’s cohesive forces allow for the property of surface tension. All of these unique properties of water are important in the chemistry of living organisms.

The pH of a solution is a measure of the concentration of hydrogen ions in the solution. A solution with a high number of hydrogen ions is acidic and has a low pH value. A solution with a high number of hydroxide ions is basic and has a high pH value. The pH scale ranges from 0 to 14, with a pH of 7 being neutral. Buffers are solutions that moderate pH changes when an acid or base is added to the buffer system. Buffers are important in biological systems because of their ability to maintain constant pH conditions.

acid: a substance that donates hydrogen ions and therefore lowers pH

adhesion: the attraction between water molecules and molecules of a different substance

base: a substance that absorbs hydrogen ions and therefore raises pH

buffer: a solution that resists a change in pH by absorbing or releasing hydrogen or hydroxide ions

cohesion: the intermolecular forces between water molecules caused by the polar nature of water; creates surface tension

evaporation: the release of water molecules from liquid water to form water vapor

hydrophilic: describes a substance that dissolves in water; water-loving

hydrophobic: describes a substance that does not dissolve in water; water-fearing

litmus paper: filter paper that has been treated with a natural water-soluble dye so it can be used as a pH indicator

pH scale: a scale ranging from 0 to 14 that measures the approximate concentration of hydrogen ions of a substance

solvent: a substance capable of dissolving another substance

surface tension: the cohesive force at the surface of a body of liquid that prevents the molecules from separating

temperature: a measure of molecular motion

Humphrey, W., Dalke, A. and Schulten, K., “VMD—Visual Molecular Dynamics”, J. Molec. Graphics , 1996, vol. 14, pp. 33-38. http://www.ks.uiuc.edu/Research/vmd/

Media Attribution

  • Figure 2.7 by Gautam Dogra
  • ice lattice by Jane Whitney
  • (b) by Carlos Ponte
  • Figure 2.10 by Cory Zanker
  • Figure 2.11 by Edward Stevens

Concepts of Biology - 1st Canadian Edition Copyright © 2015 by Charles Molnar and Jane Gair is licensed under a Creative Commons Attribution 4.0 International License , except where otherwise noted.

Share This Book

water molecule essay

U.S. flag

An official website of the United States government

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

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

  • Publications
  • Account settings

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

  • Advanced Search
  • Journal List
  • ACS AuthorChoice

Logo of acssd

How Water’s Properties Are Encoded in Its Molecular Structure and Energies

Emiliano brini.

† Laufer Center for Physical and Quantitative Biology, § Department of Physics and Astronomy, and ⊥ Department of Chemistry, Stony Brook University, Stony Brook, New York 11794, United States

Christopher J. Fennell

‡ Department of Chemistry, Oklahoma State University, Stillwater, Oklahoma 74078, United States

Marivi Fernandez-Serra

Barbara hribar-lee.

∥ Faculty of Chemistry and Chemical Technology, University of Ljubljana, Večna pot 113, SI-1000 Ljubljana, Slovenia

Miha Lukšič

Ken a. dill.

An external file that holds a picture, illustration, etc.
Object name is cr-2017-002594_0031.jpg

How are water’s material properties encoded within the structure of the water molecule? This is pertinent to understanding Earth’s living systems, its materials, its geochemistry and geophysics, and a broad spectrum of its industrial chemistry. Water has distinctive liquid and solid properties: It is highly cohesive. It has volumetric anomalies—water’s solid (ice) floats on its liquid; pressure can melt the solid rather than freezing the liquid; heating can shrink the liquid. It has more solid phases than other materials. Its supercooled liquid has divergent thermodynamic response functions. Its glassy state is neither fragile nor strong. Its component ions—hydroxide and protons—diffuse much faster than other ions. Aqueous solvation of ions or oils entails large entropies and heat capacities. We review how these properties are encoded within water’s molecular structure and energies, as understood from theories, simulations, and experiments. Like simpler liquids, water molecules are nearly spherical and interact with each other through van der Waals forces. Unlike simpler liquids, water’s orientation-dependent hydrogen bonding leads to open tetrahedral cage-like structuring that contributes to its remarkable volumetric and thermal properties.

1. Introduction: Why Is Water So Important?

An external file that holds a picture, illustration, etc.
Object name is cr-2017-002594_0034.jpg

Among materials, water holds a special prominence. It is highly abundant on Earth (occupying 1.4 × 10 9 km 3 2 − 4 ). It plays a central role in Earth’s geophysics and geochemistry, and in most of the world’s industries. It is critical for life. It occupies about half the volume inside biological cells, and it controls various important biological actions. It is a basic human need. Therefore, it is also a major source of human conflict and war. Quests for life in the universe are searches for planetary bodies that are capable of supporting liquid water. Water’s boiling and freezing points define the main temperature scales of Celsius and Fahrenheit. Some of our grandest global challenges—distributing clean water, producing cheap and clean energy, providing greater food security, green ways to produce modern chemicals, and curing diseases—depend on a better understanding of water at the molecular level. We expand on these points below.

1.1. Water Is Essential for Life

All forms of life depend on water. 5 , 6 Liquid water constitutes about half the volume of every living biological cell. 7 Water can act as a solvent, reactant, product, catalyst, chaperone, messenger, and controller. Interactions with water are a major driving force for biomolecular structure and function in living systems. They are dominant forces in the folding of proteins and nucleic acids, the partitioning of solutes across membranes, and the binding of metabolites and drugs to biomolecules. Specific water molecules often play critical roles in biological mechanisms. To better understand healthy life, and control disease, we need faster and more accurate computational classical and quantum models of water.

Life depends on the solubilities of gases in water. Humanity depends on sea life for food, and they require conditions under which oxygen (O 2 ) has sufficient solubility in water. Marine plants require carbon dioxide (CO 2 ), which must be dissolved in water, in order for photosynthesis to produce carbohydrates, which releases oxygen. Gas solubilities in water depend on temperature, pressure, and salinity.

Also important for biology are water’s surface tension and capillary action. The heights and branching of trees depends on water’s capillary action. Due to an interplay of the forces of adhesion and surface tension, water exhibits capillary action whereby water rises into a narrow tube against the force of gravity. Water adheres to the inside wall of the tube and surface tension tends to straighten the surface causing a surface rise and more water is pulled up through cohesion. The process continues as the water flows up the tube until there is enough water such that gravity balances the adhesive force. For example, when water is carried through xylem up stems in plants, the strong intermolecular attractions (cohesion) hold the water column together and adhesive properties maintain the water attachment to the xylem and prevent tension rupture caused by transpiration pull. 8

1.2. Water Is a Basic Human Need. It Is a Root of Human Conflicts

Despite water’s abundance, its distribution is increasingly problematic for the world’s growing populations; see Figure ​ Figure1 1 and Appendix A.3 .

An external file that holds a picture, illustration, etc.
Object name is cr-2017-002594_0001.jpg

Water covers 71% of Earth’s surface. Most of it is salt water. Only 2.5% of it is fresh water. And, only 1.2% of that fresh water is in rivers and lakes. The rest of Earth’s fresh water is trapped as ice in polar caps and glaciers (68.8%) or underground (30.0%).

The availability of drinking water is limited, and it is shrinking worldwide. By the year 2030, the world’s 8.5 billion people 9 will consume 6 trillion cubic meters (6000 km 3 ) of water per year. 10 While today 11% of the global population lives with poor access to clean drinking water, 11 it is estimated that in 2030 half the world’s population will be living under severe water stress. 12 It is increasingly challenging to get clean water to where it is needed. Early civilizations settled near rivers. But now, clean water is increasingly provided through water purification, desalination, 4 , 13 , 14 and transport. Therefore, clean water increasingly requires access to energy. Also, water distribution increasingly poses technical challenges, requiring advances in separating water from salts and oils at low energy costs, for example.

Water conflict is a term used to describe a clash between countries, states, or groups over access to water resources. While traditional wars have rarely been waged over water alone, 15 water conflicts date back at least to 3000 B.C. 16 The U.S. Dust Bowl drought of the 1930s, which covered nearly 80% of the United States at its peak, drove mass migration. More recent droughts occurred in the southwestern United States in the 1950s, and in California and the southern United States in just the past few years. Water has been regarded as a component of conflicts in the Middle East, 17 in Rwanda, and in the Sudanese war in Darfur. Eleven percent of the world’s population, or 783 million people, are still without access to good sources of drinking water. 11 Increased water scarcity can compound food insecurity, and put pressure on human survival.

1.3. Water Plays a Major Role in Earth’s Geophysical and Geochemical Cycles

Water plays a role in climate and weather. It is the most abundant greenhouse gas in the atmosphere, accounting for 40–70% of Earth’s retention of heat. The planet’s geochemistry is linked to the cycling of water in seas, lakes, and rivers. For example, water is transported through a hydrological cycle of evaporation, condensation, precipitation, surface and channel runoff, and subsurface flow, driven by energy from the sun. The annual flux of water through the atmosphere is about 4.6 × 10 5 km 3 /year 18 and is coupled to water’s cycling in oceans and seas between the surface and bulk.

1.4. Water Is Crucial for Industrial Processes

Almost every manufactured product uses water in at least one part of its production process. Major consumers of water are industries that produce metal, wood, food, and paper, as well as industries based on chemicals, gasoline, and oils. Water is used for fabricating, processing, washing and cleaning, diluting, cooling, or transporting products. Worldwide, agriculture and power generation are the main consumers. Agriculture accounts for 70% of all water consumption, compared to 20% for industry and 10% for domestic use. Electrical power production uses more water than any other industrial process. For example, in 2005 the United States used around 0.76 trillion L of water/day to produce electricity (excluding hydroelectric power), most of which is surface water. 19 It takes about 95 L of water to produce 1 kWh of electricity. In the food industry, about 1000 L of water is needed to produce $1’s worth of sugar.

Industrial processes are a major source of pollutants, accounting for nearly half of the water pollution in the United States. 20 Pollutants include asbestos, lead, mercury, nitrates, phosphates, sulfur, oil, and petrochemicals as well as pharmaceuticals, illicit drugs, and personal care products. 21 The world’s move toward greener chemistry—the reduction of hazardous substances in producing chemicals and reducing pollution—requires new ways to replace organic solvents with water, to better understand water’s role in reaction mechanisms, and to better understand how solutes and toxins partition into the environment. 22 − 25

Better ways are needed to separate water from other materials, such as organic solids, bacteria, and hydrocarbons. 26 Each year, 20 billion barrels of water are used in the United States to extract oil and gas. In Oklahoma, the wastewater that results from hydrocarbon extraction is so voluminous that it has been claimed to cause earthquakes. 27

2. In Some Ways, Water Is a Normal Material; In Other Ways, It Is Unusual

In certain ways, water is a fairly normal material. For example, at low temperatures, water is a solid (ice). Heating causes melting, a phase transition to the liquid state. Further heating causes boiling, a phase transition to the vapor state. Therefore, at this level, water’s pressure–temperature ( pT ) phase diagram—which expresses these general features—resembles the phase diagram of other materials. And, like other polar liquids in general, water readily dissolves salts and ions, but does not so readily dissolve nonpolar molecules, such as oils.

However, in other ways, water is complex and anomalous. Its quirkiness is reflected in various experimental observables, and requires understanding through different types of theory, simulations, and modeling. Water is more cohesive than materials made of molecules of equivalent size and shape. Water molecules associate with each other relatively tightly. Therefore, H 2 O has relatively high values of surface tension, melting point, and boiling point. And, water has density anomalies that are manifested in various ways (see Figure ​ Figure2 2 ). For example, ice floats on liquid water. In most other materials, the solid sinks in the liquid. Related to this, the slope of water’s pT equilibrium phase boundary between the solid and liquid is negative, whereas it is positive for other materials. Therefore, applying pressure melts water’s solid into a liquid, whereas applying pressure drives most liquids to freeze into a solid. And, while typical materials have only one or two solid phases, water has more than a dozen phases of its solid, ice (see also Figure ​ Figure10 10 ). Water is a polar molecule, so its liquid can dissolve polar and ionic solutes. Its thermodynamic signatures for dissolving nonpolar molecules are different from those of most other solvents. To signify that difference, it has been given its own name: the hydrophobic effect .

An external file that holds a picture, illustration, etc.
Object name is cr-2017-002594_0002.jpg

Water has volumetric anomalies. Shown here for liquid water are the temperature dependences of its heat capacity ( C P ), molar volume ( v ), thermal expansion coefficient (α), and isothermal compressibility (κ T ) for water (black lines) and a Lennard-Jones simple fluid (σ = 2.9 Å, ϵ = 0.8 kcal mol –1 ) (gray lines). Water’s heat capacity is relatively large, because water stores energy in both its van der Waals and hydrogen bonds. Water has a minimum volume (red circle), i.e., a temperature of maximum density, TMD, at 4 °C, whereas the volumes of simpler liquids increase monotonically. Cold water has a negative thermal expansion coefficient between 0 and 4 °C (red circle); heating shrinks it. Water has a negative derivative of compressibility at temperatures lower than 46 °C (red circle); heating makes it less compressible. (Data collected from from ref ( 28 ).)

An external file that holds a picture, illustration, etc.
Object name is cr-2017-002594_0010.jpg

Water’s temperature–pressure phase diagram shows its many solid phases. There are 17 known crystalline forms of water. Not pictured in this diagram are proton-ordered variations (such as ice XI which is a proton-ordered form of ice I h , where waters orient in a repeated manner rather than the more typical random fashion) and metastable forms (such as ice XVI, which is formed by solute evacuation from clathrate hydrates).

3. How Do We Know Water’s Structure–Property Relationships?

Much of what is known about water’s structure–property relationship comes from bulk measurements, such as the enthalpies, entropies, and heat capacities of pure water in its various phases; changes in those thermal quantities upon melting or boiling, or changes due to dissolving solutes; changes due to applied pressure; changes of electrical properties in applied electric fields; and measured surface tensions, for example. However, to learn how material properties are encoded within molecular structures requires more than just experiments. It also requires models, theories, and simulations. Without modeling, we cannot interpret observable properties in terms of water’s molecular structure, energetics, and population distributions. Making the structure–property connection requires knowing the driving forces. Because no single type of model currently gives a full picture, we look here through the lenses of different models and theories.

Water has been modeled variously: as a simple continuous medium for fast calculations of solvation or dielectric properties; or as simplified statistical mechanical sphere-like particles with hydrogen bonding arms; or by using fixed-charge or polarizable atomically detailed models in computer simulations; or at the (computationally expensive) quantum mechanical level for insights into the nature of electronic structure and the bonding of the atoms (see Figure ​ Figure3 3 ). Each of these approaches has its own target problems and its own research communities.

An external file that holds a picture, illustration, etc.
Object name is cr-2017-002594_0003.jpg

Different water models serve different purposes. (left) Coarse-grained and reduced-dimensionality models. These allow for the most extensive sampling of configurations, and are useful for modeling the statistical mechanics, partition functions, entropies, and heat capacities. (middle) Atomically detailed semiempirical models are used in molecular dynamics and Monte Carlo simulations of liquid and solid states. (middle left) Fixed-charge models. (middle right) Polarizable models. (right) Quantum mechanical (QM) models represent the atomic nuclei and electrons explicitly, for studying bonding. QM modeling is computationally expensive.

There are several classic reviews on the properties of water and water modeling. 29 − 33 Guillot summarized computer simulations of semiempirical models. 34 Ben-Naim has reviewed different types of models of water and solvation, including its anomalous properties. 35 , 36 A website Water Structure and Science gives an extensive collection of physical, chemical, and biological properties of water, from experiments, simulations, and theory. 37 Recently, a thematic issue on Water—The Most Anomalous Liquid was published in Chemical Reviews ( 2016 , 116 (13)) with in-depth reviews of various methodologies. And, Sun and Sun recently reviewed the role of hydrogen-bond cooperativity in water’s anomalous properties. 38 The present review goes beyond—but is is similar in spirit to—refs ( 39 and 40 ).

This review describes structure–function principles; here is a brief overview. 34 , 40

1. Water is tetrahedral. It forms hydrogen bonds. Water has strong orientational interactions in addition to van der Waals attractions and repulsions.

2. This leads to cage-like structuring, not only in the solid phases (ices) but also even in liquid water. In 1892, W. C. Röntgen (who was also the discoverer of X-rays) postulated that liquid water is a mixture of two fluids: a low-density one and a high-density one. 29 , 41 In 1933, Bernal and Fowler 42 suggested that the tetrahedral geometry of water molecules might be responsible for water’s unusual properties.

3. Liquid water is a mixture of types of structure. Over the following two decades, more microscopic models emerged, of how water’s structure gives water’s macroscopic properties. 43 , 44 From these liquid water appears to be a mixture of types of structure. Among the first quantitative models was that of Pople in 1951, 45 which went beyond treating waters as being of just two types—crystal-like or not—and supposed that waters had distributions of hydrogen bonding.

3.1. Modeling Water’s Quantum Mechanics, Electrons, and Hydrogen Bonding

The most fundamental level of water modeling is quantum mechanical. At this level, a water molecule is understood by solving the Schrödinger equation for 10 electrons in the molecule, obtaining the covalent bonding of two hydrogen atoms and one oxygen atom. The hybridization of the molecular orbitals is sp 3 , which means that a water molecule has tetrahedral structure, even though the three atoms are coplanar. This tetrahedral structure is reflected in the network-like structures in condensed phases of water (resembling those of tetrahedral elements such as Si or Ge).

The molecular orbitals (MOs) of water are shown in Figure ​ Figure4 4 . The four orbitals to the left of the dashed line each contain two electrons. To the right of the dashed line is the first empty orbital (lowest unoccupied molecular orbital, or LUMO); this plays a role in water hydrogen bonding. The lone-pair orbitals of water are a linear combination of the 1b1 (highest occupied molecular orbital, or HOMO) and 3a1 MOs. The two lone pairs of water are perpendicular to the molecular plane. The two covalent bonding orbitals are directed along the OH bond directions and are a linear combination of the first two orbitals in Figure ​ Figure4 4 . These four orbitals have maximum electron densities along the tetrahedral directions (sp 3 hybridization). A hydrogen bond (HB) forms between the positively charged hydrogen atom of the donor molecule and the lone pair electrons of the acceptor molecule.

An external file that holds a picture, illustration, etc.
Object name is cr-2017-002594_0004.jpg

Molecular orbitals of water, labeled by their symmetries.

A hydrogen bond occurs when a hydrogen atom that is bound covalently to an electronegative atom (such as oxygen, nitrogen, or fluorine) shifts its charge distribution when another nearby electronegative atom attracts it. A hydrogen bond can be regarded as an electrostatic polarization, both intramolecular and intermolecular, 46 but also entails a quantum mechanical charge-transfer component. The power of ab initio quantum mechanical modeling is in its ability to reveal the nature of bonding within water molecules in small clusters. But quantum modeling at the highest theoretical level is computationally expensive, so only very small clusters can be studied this way.

Simulations of larger structures—of sizes up to several hundred water molecules, in liquid form and in ices 47 —can be modeled by starting from an ab initio force field that gives a highly accurate quantum chemical representation of water and clusters, and then approximating the oxygen polarizability by fitting functions to gain computational speed. 48 − 54 This level of modeling shows how the shifting of the distribution of water’s electrons is needed to account for hydrogen-bond cooperativity 55 effects (whereby hydrogen bonds become stronger in the presence of other hydrogen bonds).

The properties that are captured by modeling these systems are water’s diffusion properties, its polarizability and dielectric function, 56 optical spectroscopic properties, 57 the anomalous isotope effect in ice, 58 the density anomaly, 59 IR spectra, 57 structures of small water clusters, 60 and H-bond dynamics. 61

Modeling such collections of water quantum mechanically requires simplifications and approximations. An important methodology is density functional theory (DFT) of electronic structure. Rather than to compute wave functions of all the many individual electrons and nuclei of multimolecule systems, DFT computes far fewer quantities, namely the overall spatial electron density. In the Born–Oppenheimer approximation (BO), the atomic nuclei are described by the laws of classical physics and the electronic wave functions are computed at each time step (or at each different nuclei position). In Car–Parrinello (CP) molecular dynamics , 62 the motions of electrons and ions are treated together and the electronic wave functions are propagated as classical degrees of freedom. 63 An advantage of CP is its computational efficiency.

Quantum modeling—with its ability to account for the electrons and their distributions—is a powerful way to study the nature of the bonding among water molecules, in small clusters. However, for modeling water on the larger scale—such as in its bulk liquid state, or as it undergoes phase changes to solid or vapor states, or as a solvent for ions or nonpolar molecules, or at chemical or biological surfaces—the most popular current approaches are theories and simulations at a coarser-grained or more approximate level.

3.2. Water Is Often Modeled through Semiempirical Classical Simulations Using Atomistic Potentials

Beginning around 1970, a popular approach has been to model water using semiempirical classical (i.e., not quantum) intermolecular potentials, which are sampled according to the laws of statistical physics through the use of Monte Carlo (MC) or molecular dynamics (MD) in computer simulations. In 1969, MC simulations of pure water were performed by Barker and Watts. 64 They assumed an orientation-dependent water–water pair potential consisting of a central Lennard-Jones particle surrounded by four charges, two partially positive ones centered on hydrogen atom locations and two neutralizing negative charges placed at orthogonal p-orbital locations about the oxygen center. 65 In 1971, MD simulations were performed by Rahman and Stillinger based on a symmetrical tetrad arrangement of charges centered on the Lennard-Jones site. 66 , 67 The latter evolved into the popular ST2 model for liquid water. 68 , 69 ST2 gave some of the first detailed insights into the distributions of structures in liquid water. Prior to these works, there was no way to know more than just bulk averages of experimentally observable properties, and no way to see how water’s properties were encoded in water’s molecular structure and energetics. In 1983, Stillinger noted that, because of water’s tetrahedral capacity to form multiple hydrogen bonds, liquid water is a hydrogen-bonded network that, although transient and amorphous, nevertheless has much of the ice-like character of its solid. The special directionality of the hydrogen bonds is responsible for many of the anomalous water properties. 70 He also noted that the H-bonding network in the liquid resembled the known ice network structures and he showed how the optimal structure of the water dimer is the key to understanding the strengths and properties of the hydrogen bonding. Stillinger’s computer simulations related the radial distribution functions (RDFs) of liquid water to the average number of nearest neighbors, which is close to four, because of water’s tetrahedral character, rather than the larger number of nearest neighbors that results from simpler van der Waals liquids, which pack more like marbles in a jar. Experimental measurements showed that the average number of hydrogen bonds in liquid water is in fact around 3.5 (estimates of fewer neighbors have been shown to be incorrect 71 ).

In the 1980s, simpler and computationally more efficient models emerged, such as Jorgensen’s transferable intermolecular potential (TIP) models 72 , 73 and Berendsen’s simple point charge (SPC) model. 74 − 76 The TIP and SPC models are now major workhorses for simulating water. The TIP3P model has become widely used because it is able to capture water’s liquid properties with sufficient accuracy, with only three point sites (hence, the 3P designation) of interaction, so it is computationally efficient. Because the ST2 model has five sites of point interactions, it is more costly than TIP3P. TIP3P and related models are widely used in molecular mechanics packages such as CHARMM and AMBER, as they are considered the preferred water types in the default force fields for these packages. 77 , 78 Like TIP3P, Berendsen developed the three-point SPC model of water with the same benefits of efficiency and transferability. SPC is simple, having hydrogen bond lengths of 1 Å and a tetrahedral bond angle of 109.47°. Then, in order to better capture liquid-state properties in a three-point water model, SPC/E was developed to include the missing condensed-phase electronic polarization. 74 The resulting model better reproduces pure liquid water properties including the density, diffusion constant, and liquid structure. SPC is the default model for the GROMOS package and force field. Intermediate in cost-performance trade-off are the four-point models, such as TIP4P. 79 − 82 Some water model parameters are given in the Appendix in Table 5 , and some properties are listed in Table 6 .

Some recent models are more coarse-grained. An example is the monatomic water (mW) model of Molinero and Moore, which uses a parametrized Stillinger–Weber potential. 83 Water’s hydrogen bonding is mimicked through an angle-dependent three-body potential term encouraging tetrahedral configurations. The mW model is computationaly efficient and gives the energetics, structure, and density of liquid water, as well as water’s anomalies 84 and phase transitions with comparable accuracy as most atomistic water models. The mW model was, among others, applied to studies of ice nucleation, 85 confined water, 86 hydrophobicity, 87 and clathrate hydrates. 88 A slightly more complex class of single-point models called the soft sticky dipole (SSD) model is also popular for efficient simulation of water. 89 SSD is an extension of the hard sphere BBL model, 90 and it uses point multipoles for electrostatic interactions with surrounding molecules and a tetrahedral spherical harmonic potential for hydrogen bonding between neighboring water molecules. SSD and its extended variants have been used to study liquid water structure and dynamics, 91 − 95 ice nucleation, 96 and confined water, 97 and it shows promise for general aqueous solvation investigations. 98 , 99

Even more coarse-grained are those models in which water’s atoms are collectively represented by a single interaction site. 100 − 107 Or, multiple water molecules can even be represented by a single site. 108 − 111 Yet another approach is to approximate each water molecule as being spherically symmetrical, using so-called isotropic core-softened potentials . 112 − 115 Rather than to represent hydrogen bonding as tetrahedral and dependent on orientations, these spherically symmetric models treat water–water interactions by supposing that there is tight binding at close water–water distances and weaker binding at greater water–water separations. Such models show that some volumetric anomalies can be captured without explicitly accounting for orientation-dependent hydrogen bonding.

Other recent modeling has been toward more refinement or detail, for example by incorporating three-body or many-body terms, such as in the E3B and MB-pol models. 116 − 118 It has been argued that fixed-charged models are approaching their limits of optimizability through parameter variations. 82 , 119 , 120 Further improvements beyond fixed-charge models are being sought by including polarizabilities, whereby the charge distributions within the molecule can shift depending on the molecular conformation or environment. The development of analytical potentials representing the many-body effects were recently reviewed in ref ( 121 ). So far, however, polarizable models have not yet become mainstream, partly because they are computationally expensive.

3.3. The Mercedes-Benz Coarse-Grained Model That Captures Water’s Orientational–Translational Coupling

Some of water’s major features are a consequence of its cage-like structuring and the prominent entropies and heat capacities that arise from it. Fine-grained atomically detailed simulations can be too “in the weeds” to capture these properties. Alternatively, water can be modeled using statistical mechanical liquid theories, which are good at expressing distributions of populations because of the various ways that statistical mechanics can compute averages over large ensembles. But, liquid-state theories often require potentials that have special simplicity, such as spherical symmetry, so they miss key aspects of how water’s properties arise from its tetrahedral cage-like structures.

Thus, into this breach arose the Mercedes-Benz (MB) model and its variants. It gets its name from the fact that the two-dimensional (2D) version of the MB model looks like the Mercedes-Benz logo—a circle with three radial arms that represent hydrogen bonds (see Figure ​ Figure3 3 , left). The power of MB-type models is that they do capture the translational–orientational coupling that comes from water’s orientation-dependent hydrogen bonding, while at the same time they also allow for the statistical mechanical averaging that is needed to express structural distributions, entropies, heat capacities, and free energies of large systems. 40 The Mercedes-Benz (MB) model 122 and its variants have been studied by Monte Carlo, 123 − 126 integral equations, 127 , 128 and mean-field analytics. 129 − 133 Throughout this review, we often present 2D images from this and other similar models (Rose water model 134 ) to convey structure–property relations in a simple way.

4. How Are Water’s Properties Encoded in Its Molecular Structure and Energetics?

Our starting point for understanding water’s unique properties is its tetrahedral hydrogen bonding organization (see Figure ​ Figure5 5 ). 135 − 137 In classical terms, the collective interactions between water molecules can be represented by (i) a radial dispersion attraction with steric repulsion at short-range and (ii) electrostatic interactions between spatially localized groupings of charge. In this way, hydrogen bonding is a consequence of these localized electrostatic interactions leading to nearly tetrahedral arrangement of surrounding water molecules. Water is often compared to simpler liquids, which do not structure as strongly and can be more easily treated with uniform interactions. Below, we describe structure–property relationships of water, as seen from various theoretical and modeling studies.

An external file that holds a picture, illustration, etc.
Object name is cr-2017-002594_0005.jpg

Water molecules form hydrogen bonds, giving tetrahedral structuring. The attractive interactions between water molecules can be represented with a uniform dispersion term and strong localized electrostatics, giving rise to the tetrahedral arrangement of hydrogen bonds about each water.

4.1. Water Is More Cohesive than Simpler Liquids, due to Its Hydrogen Bonding

Liquid water tends to be a more cohesive than other simple liquids, because water–water attractions arise from hydrogen bonding in addition to van der Waals interactions that are typical in simpler liquids. For example, a higher temperature is required to melt ice than to melt solids of simple liquids. And, a higher temperature is required to boil liquid water than to boil other simpler liquids. In addition, water has a relatively high surface tension, of 72.8 mN m –1 at room temperature, due to its high cohesion, the highest of the common nonionic, nonmetallic liquids.

Table 1 compares the properties of water (H 2 O) and hydrogen sulfide (H 2 S), which have similar atomic structures. Both have sp 3 hybridized orbitals, with bond angles (of HOH and HSH) being 104.45° and 92.1°, respectively. Oxygen and sulfur belong to the same group of the periodic table. But, because sulfur has twice as many electrons as oxygen, it is larger and less electronegative. Therefore, the O–H bond is much more polar than the S–H bond. Even though H 2 S has almost twice the molar mass of H 2 O, it is a gas at room temperature and pressure, while H 2 O is a liquid, indicating greater cohesion in water. Because of its hydrogen bonding, water has a higher melting point, boiling point, and heat of vaporization, as well as a higher heat capacity (which reflects the higher capability for storing thermal energy through these additional types of bonds).

propertyH OH S
molar mass [g mol ]18.01534.081
boiling point  [K]373.12212.85
melting point  [K]273.15187.45
enthalpy of vaporization [kJ mol ]40.65718.622
entropy of vaporization [J mol  K ]108.9587.9
critical temperature [K]647.1373.2
critical pressure [MPa]22.068.94
critical molar volume [cm  mol ]55.998.5
critical density [kg m ]322347
critical compressibility0.2290.284
specific heat capacity  ( ) [J mol  K ]74.53926
specific heat capacity  ( ) [J mol  K ]75.334.6

A useful way to compare the energetics of water and simpler materials is with an energy ladder; see Figure ​ Figure6 6 . States of lower energy are lower on the ladder. The laws of thermodynamics say that the densities of liquids and solids are determined, at any given temperature and pressure, by the molecular arrangement that has the lowest free energy. Figure ​ Figure6 6 , left, shows a simple material. The lowest-energy state of a simple material is the solid, typically held together by van der Waals forces. Introducing energy (say, by raising the temperature to the melting point) melts the solid, leading to fewer weaker, more disordered van der Waals interactions in the liquid state. Introducing even more energy breaks the remaining van der Waals contacts, boiling the liquid. Figure ​ Figure6 6 , right, illustrates how water is more cohesive than the simpler liquid. The melting temperature of water is higher than that for the liquid on the left in part because of the hydrogen bonding (which is tetrahedral in real water, shown as hexagonal here in a two-dimensional toy version of water). The boiling temperature of water is higher than that for the simpler material also because hydrogen bonding contributes cohesion to water. Figure ​ Figure6 6 also shows the nature of the two states of liquid water: that cold water tends to retain a little more cage-like, ice-like structure and hot water tends to retain less of it.

An external file that holds a picture, illustration, etc.
Object name is cr-2017-002594_0006.jpg

Energy–volume relationship of water, vs simpler materials. (left) Simple materials (cold) achieve low energies by tight-binding into solids, (warmer) achieve higher energies by forming looser liquid states, and (hot) achieve the highest energies when most bonds are broken in the gas phase. The black bars indicate transitions: heating melts the solid, then boils the liquid. (right) Water (very cold, ice) achieves its lowest energies through open low-density hydrogen bonded structures, (cold liquid water) achieves intermediate energies through some breakage of cages, leading to increased density , (hot liquid water) achieves higher energies by breakage of more bonds, leading to looser liquid, and (hot) achieves its highest energies, like simpler materials, by breaking most bonds, to reach the low-density gas phase.

4.2. Water’s Volumetric Anomalies Arise from a Competition between van der Waals Attractions and Hydrogen-Bond-Driven Expansion

Water has volumetric anomalies. For example, ice floats. The solids of most other materials are compact and more dense, leading them to sink to the bottom of their own liquid. For water, the solid is less dense than its liquid. This is because cold water is dominated by its hydrogen bonding, which is tetrahedral, leading to only four neighbors of any given water molecule. Cold water at low pressure tends to maximize its hydrogen bonding, which tends to cause its structure to be open and loose. Simpler materials, such as argon, pack more like hard spheres, having higher densities because each molecule has up to 12 nearest neighbors. Only under higher pressures will the open structure of ice collapse to form dense, interconnected lattices. Figure ​ Figure7 7 shows how the tetrahedral nature of the strong hydrogen bonding tends to lead to open structures, competing with the weaker and omnidirectional van der Waals interactions that simply favor more neighbors and higher densities.

An external file that holds a picture, illustration, etc.
Object name is cr-2017-002594_0007.jpg

Densities of some phases of water, illustrated in a 2D model. At high pressures, ices are dense, having some broken hydrogen bonds. At lower pressures (say, 1 bar), ices are less dense, having more optimal hydrogen bonding. Liquid water is more disordered, but still has much residual cage-like structuring.

Figure ​ Figure8 8 interprets these ladders of molecular organization in terms of the different populations, energies, and molecular volumes of the states. 40 , 129 , 130 Thermodynamics says that the densities of liquids and solids are determined, for any given temperature and pressure, by the volumetric state that has the lowest free energy. This can be expressed approximately using the energy-ladder diagrams of Figure ​ Figure8 8 . On both diagrams, the bottom energy levels show the stable states of the cold material and the top energy levels show the stable states of the hot material. Figure ​ Figure8 8 , top, shows a normal simple material: At low temperature, the material populates its low energy levels, so the material is a high-density solid. At high temperature, the material populates its high energy levels, so it is a low-density gas. In between, the material is a liquid, having intermediate density. Figure ​ Figure8 8 , bottom, shows how water is different from this. We illustrate this using five energy levels. (level 1) At low temperature below freezing, water’s energies are dominated by hydrogen bonding, so water’s structure is ice, which is low density. (level 2) Slightly warmer, above freezing, in cold liquid water, some hydrogen bonds are broken and some van der Waals interactions are made, leading to slightly denser water. (level 3) Heating still further, to warm liquid water, breaks more hydrogen bonds and makes more van der Waals contacts, so water then becomes still denser. (level 4) Heating further, to hot liquid water, now leads to breaking both hydrogen bonds and van der Waals interactions, leading to reduced density, as would be observed in a normal simpler liquid. (level 5) Higher temperatures boil the liquid, turning it into a gas, which has much lower density.

An external file that holds a picture, illustration, etc.
Object name is cr-2017-002594_0008.jpg

Water’s density (D) anomalies are correlated with shifts in Density of State (DoS) populations. (left) Same energy ladders as in Figure ​ Figure6 6 . (top right) Increasing temperature leads to shifting populations from ice-like on the left to vapor-like on the right. (bottom right) Heating a simple material drives it from solid (high density) to liquid (slightly lower density) to gas (very low density). Heating water drives it from ice (low density) to cold water (higher density) to hot water (lower density) to gas (very low density).

Some of water’s anomalies can be explained using Figure ​ Figure8 8 . First, ice floats on water because ice has lower density due to its open hydrogen-bonded tetrahedral structures. A second anomaly is that water has a temperature of maximum density (TMD, of approximately 4 °C) in its liquid range, whereas other materials are maximally dense in their solid states. Upon going from its melting point at 0 °C to its TMD at 4 °C, liquid water gets denser because heating melts some low-density, cage-structure waters into higher density van der Waals contact structures. This is also reflected in a third anomaly: just above 0 °C, water has a negative thermal expansion coefficient, indicating its increasing density with temperature. Then, heating beyond the TMD, liquid water expands like other materials do because heating loosens the intermolecular bonding.

There are practical consequences of water’s density anomalies. One is shown in Figure ​ Figure9 9 . Whereas simpler materials freeze from the bottom up, lakes filled with water freeze from the top down, since ice is less dense than liquid water.

An external file that holds a picture, illustration, etc.
Object name is cr-2017-002594_0009.jpg

Lakes freeze from the top, not the bottom. Ice floats because it is less dense than the liquid. So, even in wintry frozen lakes, fish can live at the bottom, where water is liquid.

Another anomaly is water’s nuclear isotope effect . Typically, molecules having a heavier isotope form a higher-density material than molecules having the lighter isotope. Molecules with the heavier isotope form tighter intermolecular bonds. For example, at low temperatures, 20 Ne is denser than 18 Ne. But for water, it is the opposite. That is, ice having the lighter hydrogen isotope (H 2 O) is denser than ice having the heavier deuterium isotope (D 2 O). The reason for this results from an anomaly due to a subtle quantum effect of the zero-point energies. 58 , 138 In the primary isotope effect, the covalent, intramolecular OH distance is observed to be longer than the OD distance. In the secondary isotope effect, the H-bond donor–acceptor (oxygen–oxygen) distance R changes upon isotopic substitution, being shorter for H than for D in ice. In general this occurs in materials with strong H-bonds, while in materials with weaker H-bonds the opposite effect occurs. 139 − 144 Surprisingly, the anomalous isotope effect reflected in the volume of water per molecule becomes greater at room temperature water—the volume per molecule of D 2 O is slightly larger than H 2 O, suggesting that H-bonds in H 2 O water are stronger than in D 2 O.

5. Water Has Many Solid Crystalline (Ice) Phases

The phase diagram of solid water shows many crystalline phases ( Figure ​ Figure10 10 ). The molecular structures of some of them are shown in Figure ​ Figure11 11 . The most familiar form of ice is hexagonal, ice I h . Hexagonal ice has a relatively open structure and a lower density than that of liquid water. The term hexagonal comes from its crystal structure, as shown in Figure ​ Figure12 12 , when looking at the basal plane of the crystal. This is the basal plane because the rotational symmetry in this plane gives rise to six prism planes. Water ordering and crystal growth have been shown to be enhanced on the prism planes relative to the basal planes. 145 , 146 The densities and crystal forms of various ices are given in Table 4 .

An external file that holds a picture, illustration, etc.
Object name is cr-2017-002594_0011.jpg

Different ice forms have different densities, driven by different temperatures and pressures. At the extremely high pressures on top (ice X), the red oxygen atoms and white hydrogen atoms are compressed so tightly that the hydrogen bond and covalent bond lengths in water are the same.

An external file that holds a picture, illustration, etc.
Object name is cr-2017-002594_0012.jpg

Snowflakes have sixfold symmetries because of the elementary hexagonal structures formed by hydrogen bonding in ice I h crystals. (right) Photos from Wilson Bentley taken in the winter of 1901–1902. 147

Why do snowflakes have such beautiful symmetries? Snowflake structures are a result of water’s hexagonal symmetries, which are the basis for sixfold directional growth in this plane. Snowflakes start as small six-sided plate crystals or prisms. Depending on the temperature and humidity, they may continue to grow as plates or become needles, stellar plates, or dendrites, among others, as seen in Figure ​ Figure12 12 . 147 − 149 Defects in crystal growth lead to the immense variety of snowflake shapes, captured in the historic photographs from Wilson Bentley’s early studies. 147

Some ices are proton-disordered and some are proton-ordered . This terminology refers to whether a given ice structure is achievable by multiple degenerate microscopic water configurations, or just one. In its tetrahedral lattice about the oxygen atom centers, each water molecule has six possible orientations. There is a disorder (and a corresponding entropy of R ln (3/2) 150 ) that arises from these options that are available to the molecule. Ice I h is proton-disordered because, at each lattice site, a water molecule can have different orientations. In proton-ordered ices, a water molecule can have only one configuration at each lattice site. You can experimentally craft proton-ordered forms of ice by promoting proton tunneling via the introduction of a defect (KOH can provide excess OH – ions, for example) at very cold temperatures. Then, the protons will arrange until the water molecules all order perfectly in regular directions. The entropy for these ordered forms of ice is 0, satisfying the “perfectly crystalline solid” requirement of the third law of thermodynamics. A famous (but fictional) example of a proton-ordered ice is ice-nine , made popular in Kurt Vonnegut’s novel Cat’s Cradle . 151 In that novel, it was imagined that this form of ice could nucleate whole bodies of water to freeze on contact, killing people instantly, leading to a global catastrophe. Fortunately, Vonnegut’s ice-nine is fictitious. In reality, ice IX is a proton-ordered form of ice III, and only exists at very low temperatures and high pressures, so it is not threatening to life on Earth.

6. Does Supercooled Water Have a Liquid–Liquid Critical Point?

Water can be supercooled . That is, water can be prevented from freezing—and remain a liquid—even below its normal freezing point, by careful experiments that avoid nucleation. In principle, water can be supercooled to about 150 K, which is the temperature at which devitrified (ultraviscous) water spontaneously freezes to cubic ice. 152 − 154 In practice, this point has yet to be reached because water undergoes homogeneous nucleation first. This is a kinetic point at which nucleation of the crystal happens spontaneously. The limit reached so far is about 227 K, by evaporative cooling experiments. 155 The region between the spontaneous crystallization temperature of devitrified water and the homogeneous nucleation temperature is called no-man’s land .

Two key properties of interest of supercooled water are the isothermal compressibility κ T = (∂ ln ρ/∂ ln p ) p and the isobaric heat capacity C p = T (∂ S /∂ T ) p , where p is the pressure, T is the temperature, ρ is water’s density, and S is the entropy (see Figure ​ Figure13 13 ). In 1976 Angell and Speedy showed that, as they lowered the temperature of supercooled water, these two quantities appeared to grow large, and possibly diverge, when extrapolated to −45 °C. 156

An external file that holds a picture, illustration, etc.
Object name is cr-2017-002594_0013.jpg

Supercooling water to low temperatures leads to a large divergent compressibility (top) and heat capacity (bottom). These quantities appear to diverge for temperatures approaching −45 °C. It is not clear whether this means water is approaching a typical spinodal point (i.e., the system must freeze to ice), or a point of phase transition between two liquid states. Data collected from refs ( 163 ) and ( 156 ).

What molecular structures of water can explain these apparent divergences of supercooled water? The large values of the heat capacity and compressibility indicate large variances in the energies and densities of the underlying molecular organization of water. One conjecture is that this divergence simply indicates a system reaching a spinodal point , beyond which there is no metastable liquid phase and beyond which the system must freeze. An alternative conjecture is that this divergence reflects a phase transition between two different metastable liquid phases. 157 − 159 To come to a deeper understanding, supercooled water has been studied by structural experiments. For example, X-ray structure factors have been used to determine radial distribution functions 160 and X-ray absorption and X-ray Raman scattering 161 , 162 have been used to determine first coordination shells in supercooled water.

The idea of a liquid–liquid coexistence is that there are two species of water, A and B, with temperature-dependent concentrations x A ( T ) and x B ( T ), 164 a high-density liquid (HDL) and a low-density liquid (LDL). 165 LDL could be thought of as more cage-like and HDL as less cage-like, for example. Urquidi et al. interpreted their experiments in terms of two types of dynamically interconverting microdomains having, on average, bonding characteristic of water in ice I h and ice II. 166 These hypothetical liquids derive some experimental justification from known low-density and high-density amorphous ices (LDA and HDA, respectively). 167 LDA is formed by depositing water vapor on single-crystal metallic surfaces, and HDA is obtained applying pressure to ice I h . The idea of a liquid–liquid critical point is partly motivated by the observation of a first-order phase transition between LDA and HDA.

The properties of supercooled water have been explored by computer simulations. Supercooled water likely has a critical point below the homogeneous nucleation temperature, so it cannot be probed directly by experiments, hence the need for computer simulations of liquid water in no-man’s land . 168 − 170 Definitive results have been difficult to obtain because (1) phase equilibria are slow processes that challenge computational resources and (2) any possible free energy barriers are small and subtle, requiring extensive computational sampling, and different models can give different results. 162 , 171 − 177 At the center of this animated debate has been the ST2 water model. 178 Recent modeling, however, does appear to support the view that supercooled water undergoes a liquid–liquid phase transition. 179 − 181 Further support comes from a study in 2015 by Smallenburg and Sciortino, who proved that tetravalent model systems similar to water have two stable supercooled liquid phases. 182

7. How Is Water Structured around Solutes That Are Nonpolar?

Water is called the universal solvent because it dissolves a wide variety of substances. Water is polar, so it readily dissolves charged or polar solutes. Water also dissolves some molecules that have nonpolar character, such as aromatics and surfactants. However, water is not a good solvent for nonpolar molecules such as hydrocarbons (oils). This is the basis for the expression that “oil and water don’t mix”. This avoidance tendency of oils for water is the basis for many important processes, such as the following. Surfactants and soaps spontaneously form micelles in water; it is the basis for their cleaning actions. Lipids spontaneously form bilayers, forming the structures of cell membranes, which defines the “self” of the cell. Drugs and metabolites either partition into lipid bilayers or not, depending on their degree of nonpolarity, dictating whether or not they have biological or medicinal properties. Protein molecules, which tend to contain about equal numbers of polar and nonpolar amino acids, fold into compact structures—the nonpolar parts forming a core, minimizing their contact with water. These folded structures perform most of the chemical and mechanical properties of living cells. Many of the binding processes in biology—of proteins to other proteins or to DNA or to drugs and metabolites—are driven predominantly by the degree to which nonpolar molecules tend to avoid solvation in water. Many of the methods of separation or analysis of analytical chemistry, such as reversed-phase chromatography, are governed by the varying affinities of solutes for water. Toxins and pollutants partition into environments, including into otherwise clean water, depending on their degrees of nonpolarity.

You can study solutes dissolved in water using partitioning experiments . These experiments measure the solute concentration in water and the solute concentration in an oil or vapor phase that is in equilibrium with the water. Polar solutes tend to concentrate more in water, whereas nonpolar solutes tend to concentrate more in oil phases. The space of all possible solute molecules is huge. Therefore, the goal of determining solvation in water of a broad range of solutes has been made simpler using the companion ideas of model-compound partitioning and additivity relationships . The idea is that an arbitrary solute can be thought of as a collection of smaller component substituent moieties, such as individual methylene or carbonyl or alcohol groups. The free energy of partitioning of these components can be measured in a partitioning experiment. The free energies of solvation of all the moieties can be added together to compute the solvation free energy for the whole solute molecule. For example, by measuring the concentrations in water of a series of alcohols—methanol, ethanol, propanol, butanol, etc.—Tanford determined that the free energy of transferring each individual methylene group into water was 825 cal mol –1 . 183 That number has then been used to estimate the free energy of solvation of methylene groups in arbitrary solutes, as an additive term for each such group. 184

The promise of such model-compound studies has been twofold. First, it has offered the prospect that, by simple experiments on a small number of small molecules, many chemical processes could be predicted on a much larger set of molecules, including the folding of proteins, formation of lipid bilayers, binding of drugs to proteins, partitioning of toxins in the environment, and the transport of drugs through cell membranes, for example. Second, model compound experiments could give insights into how water is structured around solutes.

The idea of hydrophobicity scales arose in the 1970s from these model-compound-plus-additivity studies. Correspondingly, a compound was said to lie along a single axis—from polar to nonpolar. The notion was that a solute’s partitioning into one hydrocarbon would be the same as into any other hydrocarbon (alkanes, alkanols, or cyclohexane, or octanol, for example), irrespective of the chemical details of the particular hydrocarbon or solute. However, it is now clear that nonpolarity is not independent of chemical details, and that hydrophobicity scales differ from one another, so the notion of “hydrophobicity scale” is more qualitative than quantitative. 185

Similarly, model-compound studies have been found to be more correlational than quantitatively predictive, due to nonadditivities . Figure ​ Figure14 14 shows an example of a nonadditivity in free energies of transfer of xylenol solutes into water. The methyl groups of 2,6-xylenol ( Figure ​ Figure14 14 , right) sterically crowd a first-shell water molecule, leading to poorer solvation. In short, perturbing this single water molecule affects the whole solvation cage, causing a loss of 1.4 kcal mol –1 of solvation free energy, out of a total solvation free energy of −6.3 kcal mol –1 in the mostly unperturbed 3,5-xylenol ( Figure ​ Figure14 14 , left). 186 Ordinarily, it would be expected that if one solvent molecule were perturbed, say, out of 20 solvent molecules in a first solvation shell, then a solvation free energy of −6.3 kcal mol –1 would have been affected by only about 1/20 = 0.3 kcal mol –1 ; see 2,4-xylenol ( Figure ​ Figure14 14 , middle). But when the solvent is water, very small perturbations that lead to changes in even one hydrogen bond in one water molecule can have outsized energetic consequences.

An external file that holds a picture, illustration, etc.
Object name is cr-2017-002594_0014.jpg

Solvation free energies can be nonadditive because the solute can perturb waters away from their cage-like favored structures. In a series of xylenol molecules, the different methyl group arrangements perturb the water solvation of the (red) hydroxyl group of each solute, particularly in the first solvation shell. The (blue) water occupancy from molecular simulations shows how displacing water from the ideally H-bonded structure leads to a decrease in the Δ G solv (here in units of kcal mol –1 ).

Insights about solvation structures can be obtained by decomposing solvation thermodynamics into its two components ( Figure ​ Figure15 15 ): (i) a nontransient cavity forms in the solvent and then (ii) the solute enters and interacts with the solvent cavity. 29 , 187 − 190 The first step describes the reversible work spent on the water–water interactions to open a cavity of appropriate size and shape ( reorganization free energy). The second step expresses the free energy of the solute–water interactions ( binding free energy). In general, the driving forces of the solvation process can be subtle and complex, 29 , 187 − 197 but we give a general overview below.

An external file that holds a picture, illustration, etc.
Object name is cr-2017-002594_0015.jpg

The solvation process can be decomposed into two steps. (left) Prior to solvation, the solute (gray) is in the vapor phase (white). (middle) A cavity opens in the water solvent (blue). (right) The solute inserts into the cavity.

7.1. Oil and Water Do Not Always Mix: The Hydrophobic Effect

A longstanding rule-of-thumb about mixtures is that “like dissolves like”. In general, when two species A and B are combined in a mixture, the AA and BB attractions tend to be stronger than the AB attractions. The fact that oil and water often do not mix is consistent with this rule. But, there is an important difference between water and simpler systems. In simpler systems, immiscibility is because of energies. For oil and water, the thermodynamic signature of the immiscibility (at room temperature) is entropic. This is manifested in the different temperature dependences of solvation; see Figure ​ Figure16 16 . This figure shows how the entropy, enthalpy, and free energy depend on temperature when dissolving toluene in water. Interestingly, even though the entropy and enthalpy of aqueous solvation of nonpolar solutes change substantially with temperature, the solvation free energy is relatively independent of temperature.

An external file that holds a picture, illustration, etc.
Object name is cr-2017-002594_0016.jpg

Dissolving oil is entropically favorable in hot water, but entropically unfavorable in cold water. For toluene in water. 198 The solvation free energy depends little on temperature. The enthalpy and entropy depend more strongly on temperature, and they compensate. In cold water, the solute induces more ordering than in bulk waters, and the cages have good H-bond and solute interactions. 199 In hot water, insertion of oil breaks potential water–water hydrogen bonds, leading to both higher entropies and higher enthalpies.

The structural explanation of aqueous solvation dates back at least 70 years. 200 , 201 Liquid water can be viewed as an ensemble of cage-like structures, not a single ice-like structure. 202 Cold water has more population of open-cage states. 203 Introducing a small hydrophobic particle shifts the equilibrium further toward these more open and ordered states, 187 and water has slower dynamics in the first few solvation shells. 204 , 205 Hotter water has less cage-like organization. 40 , 206 − 208 Introducing a solute into hot water acts more like traditional “like-dissolves-like” situations, where the solute insertion into water is unfavorable for energetic reasons.

The solubilities of nonpolar molecules depend on their sizes. Consider a series of nonpolar solutes in water, having increasing radii. They will induce first-shell water structuring that differs depending on the solute size; see Figure ​ Figure17 17 .

An external file that holds a picture, illustration, etc.
Object name is cr-2017-002594_0017.jpg

A small solute is compatible with water’s natural cages. A large solute does not fit in a cage. In cold water, small solutes can fit in the available cavities with minimal perturbation of the water structure. This process is favored by enthalpy and opposed by entropy. However, in cold water, big solutes do not fit in preexisting cavities. First-shell solvating water molecules around large nonpolar solutes are more disordered. Dissolving large solutes in cold water is opposed by enthalpy (breaks hydrogen bonds) and favored by entropy.

In short, a small nonpolar solute (say, the size of xenon) does not dissolve well in water at room temperature for entropy reasons. Small solutes can fit into natural hydrogen bonded water cages, but they distort those cages, which is entropically unfavorable. 40 , 207 , 209 , 210 In contrast, a large nonpolar solute (say, a sphere bigger than 1 nm diameter) does not dissolve well in water at room temperature for enthalpy reasons. A large solute cannot fit within water’s common small hydrogen-bonding cages. A large solute must enter into a larger unstructured volume of water, where hydrogen bonding is largely already broken. The different physics of the transfer of a small big hydrophobic object in water is reflected also in the magnitude of the transfer solvation free energy per unit surface area. For small solutes this quantity is about 25 cal mol –1 Å –2 . This reflects the cost of ordering water molecules around the solute. For big solutes the solvation free energy per unit area tends to the value of the water macroscopic surface tension, which is about 75 cal mol –1 Å –2 . This reflects the cost of breaking the HB network around the solute. 40

The solvation free energy of a hydrophobe can also be understood in terms of the probability of finding a cavity in water of the appropriate size. Information theory links those two quantities. 211 This way of considering hydrophobic solvation is of particular interest in biology in order to understand the wetting–dewetting of biological surfaces. 212

7.2. How Do Two Hydrophobes Interact in Water?

The potential of mean force (PMF) is the free energy of bringing two particles together in a solvent from a large original distance apart to a separation r from each other. Figure ​ Figure18 18 shows the PMF of two hydrophobic spheres in water. The curve has some minima and maxima. The first minimum ( contact minimum ) represents the free energy of the two particles brought into direct contact. This configuration is favorable because the direct-contact state minimizes the total water-accessible surface of the two hydrophobes, relative to all other separations. The particles cannot approach closer than this due to steric repulsion. The second minimum ( solvent-separated minimum ) is favorable because then the spheres fit compatibly within water’s caging structure. 211 , 213 − 220 The unique aspect of water here is its ability to form cage structures, causing the solvent-separated state to be relatively stable for some hydrophobes.

An external file that holds a picture, illustration, etc.
Object name is cr-2017-002594_0018.jpg

Water is structured differently around two hydrophobes at different separations. The PMF is the reversible work spent to bring two hydrophobic particles from infinite distance to the distance r . The leftmost minimum shows that two hydrophobes in contact are stable in water. The middle minimum (solvent-separated state) shows partial stability when two hydrophobes are both in water cages, separated by a layer of water. The maxima (unstable states) are hydrophobe separations that have unfavorable water configurations.

7.3. Water Pulls Away from Hydrophobic Surfaces

Water molecules tend to avoid hydrophobic surfaces. Waters do not form hydrogen bonds with such surfaces. 209 , 221 Even more unfavorable is when water becomes squeezed between two hydrophobic surfaces. Such confinements can be sufficiently unfavorable that water molecules between nonpolar planes will vaporize inside, and escape the confinement. This has been called a drying transition or dewetting . A more subtle consequence is that water at a hydrophobic surface will have larger density fluctuations than it will have in the bulk. 222 − 229 Lum, Chandler, and Weeks noted that this repulsion between a hydrophobic solute surface and surrounding waters depends on the size of the solute: bigger flatter surfaces tend to exclude water more strongly. 221 Therefore, large hydrophobic objects will tend to cluster together, to squeeze out the water molecules in between them. 209 , 227 , 228 , 230 − 241 MD simulations show that (1) while dewetting does happen in the melittin tetramer protein, 225 (2) it does not happen in the two-domain protein BphC, 223 , 224 and (3) the dewetting phenomenon can disappear in the presence of small additional interactions, and depends on force field parameters. 226 While water may not fully deplete near hydrophobic surfaces, it may have larger fluctuations than it has in the bulk.

8. Water Forms Solvation Structures around Ions

8.1. ions are kosmotropes or chaotropes, depending on whether they order or disorder waters.

Put an ion in water. In the ion’s first solvation shell, water molecules will adopt different types of structure depending on the size and charge of the ion. 242 Gurney defined some ions as being structure makers or kosmotropes , and other ions as being structure breakers or chaotropes . 243 Water’s structure-making or structure-breaking tendencies are defined by experiments that show how each type of ion changes the viscosity of water, or its entropy of solvation in water, for example. 126 , 242 If a type of salt ion is added to water that has a negative Jones–Dole viscosity B coefficient or has a negative entropy of solvation, that type of ion is called a structure maker. 244

In 1957, Samoilov 245 also proposed water structuring around ions is reflected in a dynamical property, namely the activation energy required to strip a water molecule away from the first solvation shell of an ion. A first-shell water molecule around a kosmotropic ion is more tightly bound to the ion than that water is bound to another water. He referred to this as a positive activation energy. In contrast, a first-shell water molecule around a chaotropic ion is bound more weakly to the ion than that water is bound to another water. He called this a negative activation energy. 245 The structures and dynamics of water in ion hydration shells have been studied extensively by diffraction and spectroscopic measurements, as well as by computer simulations (for reviews, see refs ( 242 and 246 − 248 ).).

Figure ​ Figure19 19 shows that small ions cause electrostatic ordering of solvating waters, while large ions cause hydrophobic ordering (shown in the Mercedes-Benz-plus-dipole model 40 , 126 ). A small ion’s charge can come close to a water molecule, resulting in a strong electrostatic attraction for the water’s dipole. This is the nature of kosmotropic ordering of water around small ions such as Li + and F – . In contrast, the charge at the center of a large ion cannot come close to a water molecule, so the electrostatic interactions with water are weak. Around large ions, water molecules form hydrogen bonds with other water molecules, as they would do around nonpolar solutes. Chaotropic ordering of water around large ions such as Cs + and I – resembles hydrophobic water structuring. 40 , 126

An external file that holds a picture, illustration, etc.
Object name is cr-2017-002594_0019.jpg

Around small ions, waters become electrostatically ordered . Around large ions, waters become hydrophobically ordered . The coloring shows the probability density of water dipole positive (blue) and negative (red) charges as well as water–water hydrogen bonding arms (orange). About small ions, water is highly electrostricted.

An anion does not have the same effects as a cation of the same size. The negative charge on water’s dipole is at the center of the water molecule. The positive charge on water’s dipole is near the outside of the water molecule. Therefore, an anion can come closer and interact more strongly with a water’s dipole than a cation can. This leads to a notable difference between anion and cation size required to achieve a given level of water ordering. 126 , 249 − 251

8.2. In the Hofmeister Effect , Salts Can Drive Nonpolar Molecules To Aggregate or Disaggregate in Water

At very low concentrations, nonpolar solutes will dissolve in water. Increasing its concentration causes a solute to reach its solubility limit, and to then aggregate. Now, to these hydrophobe–water solutions, add salt. Some types of salts will increase the solubilities of nonpolar solutes (called salting in ), and other types of salts will decrease the solubilities (called salting out ). 252 − 254 These effects of salts on hydrophobe solvation were first discovered by Hofmeister, 255 , 256 and are widely applied to dissolving or precipitating proteins, which are partly composed of hydrophobic amino acids. The Hofmeister effect is commonly modeled by the Setschenow equation: 254 ln( c i / c i (0)) = − k s c s , where c i and c i (0) are the solubilities of the hydrophobe in a solution of salt and water, respectively, and c s is the concentration of the salt. k s is the salt’s Setschenow coefficient; it depends on the type of salt, as well as on the nature of the hydrophobic solute. 254 , 257 − 260 At small salt concentrations, the cation and anion effects on the hydrophobe solubility are typically independent and additive. 261 , 262 The Hofmeister series is a list in which different types of ions are rank-ordered in terms of how strongly they modulate hydrophobicity.

Small ions tend to cause salting out: adding salt reduces hydrophobic solubilities in water. Large complex ions tend to cause salting in: adding salt increases nonpolar solubilities in water. Molecular dynamics simulations of Smith 263 and Kalra et al. 264 indicate that salting out is because the hydrophobe is excluded from the first hydration shell of the ion. Salting in is because the hydrophobe can occupy the ion’s solvation shell. And, like chaotrope/kosmotrope properties, Hofmeister effects tend to correlate with the charge densities on the ions.

Figure ​ Figure20 20 illustrates the structural basis for Hofmeister effects. 40 , 126 Small ions exclude hydrophobes because small ions bind water molecules quite tightly. This exclusion increases the hydrophobe concentration in the remaining volume of the solution, driving the hydrophobes to aggregate with increasing ion concentration. Large ions do not exclude hydrophobes because large ions do not bind waters so tightly. Therefore, adding large ions to solution does not drive increased hydrophobe concentration in the remaining solution volume. The black bars in Figure ​ Figure20 20 show the hydrophobe exclusion volume of different ions, as well as the well-known observation that the smaller ions tend to have the larger exclusion volumes. 265

An external file that holds a picture, illustration, etc.
Object name is cr-2017-002594_0020.jpg

Hydrophobes are excluded around small ions. Hydrophobic solutes (gold) insert more readily into the inner solvation shells of large ions than small ions. (orange) Hydrophobe density. (light gray lines) First and second solvation shells (black and gray bars underneath) show the effective size of the ion, from the perspective of external solutes.

8.3. When Two Ions Interact in Water, Both Solvation Shells Determine the Solution Properties

Figure ​ Figure21 21 shows the PMF of two ions coming together in water. It shows how the solvation-shell waters are structured at different ion separations. When the ions are far apart, each ion’s solvation shell is structured as described above for the isolated ions. When the two ions are separated by about one layer of water, the bridging waters between them will be structured by multiple interactions. Each bridging water interacts with other bridging waters through hydrogen bonding, and each bridging water interacts with each ion through its water dipole. When the two mobile ions come into contact, the ion–ion electrostatics can also contribute substantially to the free energy. Extensive computer simulations using different water models show that the shape of the PMF depends on all these factors. The resulting free energy from the sum of the factors can be quite different for ions of different sizes and shapes. 266

An external file that holds a picture, illustration, etc.
Object name is cr-2017-002594_0021.jpg

The water structure around an ion pair depends on the cation–anion distance. At large separations, each ion has its own solvation shell. At intermediate separations, the ion pair is stabilized by bridging waters . Ion–ion contacts of opposite charge are stabilized by electrostatic attractions, in addition to the water forces. (blue) Positive charge density of the waters. (red) Negative charge density of the waters. (orange) Density of the water–water hydrogen bonding arms.

There is an interesting puzzle of ion pairing. Some salts are more soluble in water than others. When a salt is composed of a small anion and a small cation, say LiF, it is relatively insoluble in water. When a salt is composed of a big anion and big cation, say CsI, it is also relatively insoluble in water. But, when a salt is composed of a small ion and a large ion, say CsF, then it is relatively soluble. Collins explained this through his law of matching water affinities . 251

The structural basis for Collins’ law is shown in Figure ​ Figure22 22 . Two small ions stick together (their contact state is most stable) because their ion–ion charge attractions dominate the energetics. Two large ions stick together because they act like hydrophobes (since their charge interactions are weak, because the two ions cannot come sterically close together). But the middle diagram in Figure ​ Figure22 22 shows that small–large ion pairs tend to be most stable in their solvent-separated states, hence these are the least “sticky” of the salt types. The reason this state is stable is because the small ion attracts a water cage around it for electrostatic reasons, and the large ion is compatible with a water cage for hydrophobic reasons. 267 When ions are stable in water-separated states, they will also tend to dissolve well in water.

An external file that holds a picture, illustration, etc.
Object name is cr-2017-002594_0022.jpg

The water structure around an ion pair depends on the size of each ion. Water is more electrostricted around small ions. 266 The ion–ion contact state is most stable for small +/small – , because of electrostatic attractions. The contact state is also most stable for large +/large – , because of hydrophobic-like water structuring. The most stable state for large-ion/small-ion pairs is solvent separated.

Therefore, the properties of aqueous solutions of even the simplest ions result from a subtle balance of geometry, hydrogen bonding, and charge interactions. When additional forces are also involved, such as when ions are near curved surfaces or protein binding sites, it can further tip that balance. 268 On the other hand, aqueous solvation of solutes that are polar but not ionic can be simpler. 269 , 270 For example, around alcohols, water–water hydrogen bonds can be replaced by water–alcohol hydrogen bonds, but much of the rest of the water structuring is unaffected. In these cases, charge and geometry are less complicating factors.

9. Water in Confined Spaces and at Liquid–Vapor Interfaces

Water is commonly in contact with surfaces. Examples include water permeating through granular or porous or supramolecular structures or gels, 271 or inside crowded biological cells, or bound to proteins or DNA, or at interfaces with air or oils. 272 Surfaces can constrain or induce water structuring, 273 by promoting or interfering with water–water hydrogen bonding. For example, since water cannot form hydrogen bonds to hydrophobic surfaces, water tends to move away from them toward locations most compatible with forming water–water hydrogen bonding instead. 272 , 274 An example of a hydrophobic surface is the air–water interface. At surfaces with air, water can lose hydrogen bonds simply because of the geometric constraint imposed by the surface; 272 , 275 see Figure ​ Figure23 23 . Some recent experiments suggest that the lost hydrogen bond is due to the OH group, 272 , 275 while others suggest that it is the lone pair on the oxygen pointing away from water’s surface. 276 − 278 Air–water interfaces are also slightly enriched in hydronium ions (H 3 O + ), but not hydroxide (OH – ). 272 , 274 , 279 Hydronium ions are only weak H-bond acceptors, so they cannot compete with water–water hydrogen bonds in the bulk liquid, and therefore hydronium ions tend instead to concentrate at surfaces.

An external file that holds a picture, illustration, etc.
Object name is cr-2017-002594_0023.jpg

Water molecules orient at interfaces to favor hydrogen bonding. Air above liquid water acts like a hydrophobic surface. In order to minimize the loss of H-bonding interactions, interfacial water will tend to orient such that either a single proton or electron lone pair points toward the air. 275

Hydrophobes in water tend to concentrate at air–water interfaces. This is because hydrophobes tend to localize wherever they are able to break the fewest water–water hydrogen bonds. Surfaces have a lower density of water–water hydrogen bonds than bulk water has. 272 Acids and bases can also differentially localize at surfaces, depending on the preferences of the protonated or deprotonated forms of a molecule. For example, consider acetic acid in a solution with water at a bulk pH equal to the p K a (i.e., 4.8). In the bulk, there will be equal concentrations of the protonated and deprotonated forms. But at water’s surface, acetic acid is found predominantly in its protonated form. 274 At pH ≪ p K a , both the bulk and the surface of water are enriched in the protonated species. At pH ≫ p K a the deprotonated species dominates the bulk, and the surface is depleted of both forms. 272 Other types of ions, too, are enriched or depleted at water surfaces depending on the preferential interactions of those ions with surface vs bulk waters. 272 , 280 − 283

9.1. Water Is Structured in Nanotubes Partly by Hydrogen Bonding

Water inside nanotubes can be sterically constrained; 271 see Figure ​ Figure24 24 . Water’s organization inside such confined spaces can depend on the size and shape of the cavity. In narrow hydrophobic pores ( Figure ​ Figure24 24 , left), waters can form hydrogen bonded chains, where each water hydrogen bonds to one water neighbor in front and one behind. According to semiempirical simulations by Hummer et al., such water wires can explain how water is transported up to 3 orders of magnitude faster in nanotubes than in the bulk. 284 , 285 Water’s normal slow mobility in the bulk is because it moves by breaking and making hydrogen bonds to neighboring waters. Its high mobility in nanotubes is because water does not break or make hydrogen bonds to the nanotube walls as it flows. However, other studies show the opposite; namely that water flow is retarded in certain kind of nanotubes. 286 , 287

An external file that holds a picture, illustration, etc.
Object name is cr-2017-002594_0024.jpg

Nanoconfined water is geometrically constrained (left) in a tube of diameter 8 Å and (right) in a tube of 11 Å. The hydrogen bonding between two waters or between water and the surface depends on the size and shape of the confinement. 271

Putting water into some physical confinement can shift water’s pT phase diagram. Different confining surfaces can either raise 288 or lower water’s melting point. 289 Smaller nanotubes tend to lead to higher boiling points and depressed freezing points of water. 271 Also different forms of ice may form within confined spaces. Depending on the diameter, as well as their interior surface characteristics, nanotubes may exhibit ion selectivity, similar to that of the ion channels, which is attributed mostly to the formation of ion hydration shells. 290 , 291

9.2. Icy Frontiers: Ice Changes under Confinement and in Clathrate Cages around Hydrocarbons

Several of water’s ices are observed under pressures that are greater than 100 atm and are applied omnidirectionally. What happens if water is squeezed in only a single direction, such as between two smooth plates? Such a situation has been studied for ice that contracts upon freezing in a bilayer form, sometimes called Nebraska ice because “Nebraska” is a Native American term meaning “flat water” and because the modeling was performed at the University of Nebraska. 292 Modeling has explored the stabilities of known ice polymorphs. 293 , 294 Computer simulations have predicted phases of ice not yet seen experimentally, 96 , 295 and have explored the structuring of ice that is confined within nanotubes, 296 − 299 ranging from linear chains, to helical H-bonded spirals, to stacked pentagonal rings, and more. 297 These repeating patterns are similar to the optimal water cluster arrangements for small numbers of water molecules, 300 apparently consistent with the principle that water adopts structures having maximal hydrogen bonding subject to the given geometric constraints.

Sometimes ice itself is the constraining vessel. Other molecules can be captured within ices, in the form of clathrate hydrates . Clathrate hydrates are ice cages that encapsulate small molecules, typically hydrocarbons. There are several regular clathrate structures, and Figure ​ Figure25 25 shows the two most common forms. The shape and size of the caged solutes, as well as the pressure and temperature, will direct the type of clathrate structure adopted by the surrounding water.

An external file that holds a picture, illustration, etc.
Object name is cr-2017-002594_0025.jpg

Clathrate ice structures can act as cages for gas molecules. Two of the most common clathrates are the cubic structure type I (CS-I or sI) and cubic structure type II (CS-II or sII). The imaged clathrates are oriented to highlight CS-I’s simple cubic and CS-II’s face-centered-cubic lattice structures, and respective cage cutaways with space-filling CO 2 and CH 4 guests are shown below.

Among the most notable clathrate hydrates is methane clathrate. 301 These are found naturally in large quantities on the cold ocean floor. Clathrate hydrates can be problematic commercially because they limit gas and oil extraction by clogging up the transport of hydrocarbons through pipelines. Despite this troublesome aspect, their large natural abundance and high methane density mean methane clathrates represent an enormous potential energy source. 302 However, extraction is problematic and is at present not economically competitive with standard drilling techniques.

Under the proper conditions, clathrate hydrates can be grown, often by doping ice using a hydrogen-bonding solute that can catalyze hydrate formation through the uptake of surrounding gas solutes. 303 − 306 They are promising for gas storage and transport, such as for sequestering CO 2 . Growing such clathrate hydrates also provides an avenue for finding new structures of pure-water ices. Solutes can be extracted from clathrate hydrates by vacuum evacution to empty the ice cages, leading to new potential phases of pure ice. 307 , 308 Such a strategy was recently applied by Falenty, Hansen, and Kuhs to evacuate Ne gas from a Ne CS-II clathrate hydrate. 309 The resulting ice XVI structure is the lowest-density ice polymorph that is observed experimentally. Not unexpectedly, it is more delicate than the unevacuated Ne clathrate, collapsing at temperatures above 145 K.

10. Molecular Structure Governs Water’s Dynamical Properties

10.1. water diffuses more rapidly above its glass transition temperature than below it.

By studying water’s diffusion and viscosity properties, it is found that water is a complex glassy material. Cooling a liquid slowly causes it to freeze to a stable crystalline state. But cooling a liquid rapidly can lead to kinetically trapped states that are glassy , i.e., which do not have regular order and in which transport is sluggish. 310 Water undergoes such a glass transition to an amorphous solid. In water, the characteristic relaxation times become of the order of 100 s and the rate of change of the volume and entropy decreases abruptly (but continuously) to that of a crystalline solid. 311 Glassy water dynamics can be explored using computer simulations of supercooled water. But, such simulations have been challenging because the time scales are very slow—15 orders of magnitude slower than in normal liquid water. 312 Glassy water is studied either through the dynamics of the molecules and HB network directly 313 − 316 or by relating the dynamics to equilibrium properties. 312 , 317 , 318

But, even as glassy materials go, water is unusual. A key diagnostic is the rate of change of a material’s viscosity with temperature as the temperature approaches its glass transition point. 319 − 321 Plotting the logarithm of the viscosity of a liquid vs the inverse of the temperature of the liquid, Angell defines strong glass formers as those having straight lines on this Arrhenius-type figure. He defines fragile glass formers as having curved lines, where the viscosity is insensitive to temperature in the hot liquid and strongly dependent on temperature in the cold liquid. Strong glass formers are regarded as having memory of their molecular arrangements above their glass transition. Fragile liquids forget their glass structure much faster upon crossing the glass transition into the liquid phase. Water is more complicated than either of these behaviors, because of the complexity of structure in the supercooled liquid state. The high-density and low-density components of water each have different viscosity characteristics.

10.2. Cold and Supercooled Water Diffusion and Viscosity Depend on the Relative Population of High and Low Density Water

In simple liquids (and in hot water), increasing the pressure increases the viscosity and decreases the diffusivity because applying pressure leads to crowding of the molecules, making their motion more sluggish. However, cold water behaves differently. At temperatures below 306 K the viscosity of water decreases with increasing pressure. 322 , 323 Below 283 K the diffusivity of water molecules increases upon increasing the pressure of the system. 324 Anomalous behavior is also observed in the sound velocity in cold water. 325 What explains these results? According to Le Chatelier’s principle, applying pressure squeezes a system into a denser state. In the case of cold water, applying pressure shifts water structures, crunching cage-like waters into van der Waals clusters. This breakage of hydrogen bonding frees up waters to move faster. 163 , 167 , 323 , 326 − 329

10.3. Protons and Hydroxide Ions Diffuse Rapidly in Water

Small molecules and ions diffuse through liquids. Their diffusion rates typically depend on the radius of the diffusant and the viscosity of the solvent. However, when the liquid is water, there is a remarkable exception. Water’s own component breakdown products—its protons and hydroxide ions—diffuse much faster than other ions do in water. The mobility of H + is 7 times the mobility of Na + , and the mobility of OH – is 2.5 times the mobility of Cl – ; see Table 3 . This implies that there is a distinctive transport mechanism for H + and OH – compared to other ions in water. Ionic diffusion in water plays crucial roles in processes in biology (e.g., proton transfer in enzymes and biological channels), in industry (e.g., transfer of ions through fuel cell membranes), and in environmental processes (e.g., ion transfer on ice surfaces facilitating atmospheric reactions). Water ionization is a key component of aqueous acid–base chemistry.

Water’s autoionization , H 2 O ⇌ H + + OH – , is a rare event. It occurs on the ∼10 h time scale. 330 In aqueous solutions, the proton (H + ) does not exist in unhydrated form. The proton in water forms a hydronium ion , H + (H 2 O) or H 3 O + , which is itself short-lived. 331 The hydronium ion can donate three hydrogen bonds. Hydronium makes strong hydrogen bonds, 332 which can influence more than 100 surrounding waters, 333 and can form hydronium chains. Manfred Eigen suggested that the proton is localized on an individual water molecule, and that the H 3 O + gets further solvated by three water molecules, i.e. H 3 O + (H 2 O) 3 or H 9 O 4 + ; this is called the Eigen cation ; 334 , 335 see Figure ​ Figure26 26 . Alternatively, Georg Zundel suggested that H + is shared equally between two water molecules, forming the (H 2 O)H + (H 2 O) or H 5 O 2 + cation, called the Zundel cation ; 336 see Figure ​ Figure26 26 . Other types of structures for the hydrated proton have also been proposed, 337 − 339 as well as for the hydrated hydroxide ion (OH – (H 2 O) n ). 274 , 340 − 345

An external file that holds a picture, illustration, etc.
Object name is cr-2017-002594_0026.jpg

In liquid water, protons are hydrated. (A) On the Eigen cation , the proton H + is localized on one water molecule. The resulting H 3 O + is hydrogen bonded to three surrounding water molecules. (B) On the Zundel cation , the H + is shared between two water molecules.

Simple diffusion does not explain the anomalously high electric mobilities of H + and OH – . A mechanism was proposed in 1806 by Theodor von Grotthuss; 346 , 347 see Figure ​ Figure27 27 . Protons hop along a hydrogen bond network of neighboring water molecules, like a “bucket brigade” by cleaving and forming covalent bonds. Recent computer simulations give additional insights. First, it is found that the Eigen and Zundel cations are only limiting ideal structures, and that there are delocalized defects, giving a broader distribution of structures. 341 , 348 , 349 Second, the proton transfer in water is not exclusively by stepwise hopping, but includes broad pathways, time scale distributions, modes of transfer, and concerted motion of multiple atoms. 350 This migration of charge involves bursts of activity along proton wires . 351 − 353 A proton can jump 4–8 Å across several water molecules on the subpicosecond time scale, followed by resting spells. 352 Third, it is found that the transition state in proton transfer is the formation of special water pairs having rare ultrashort hydrogen bonds, which can lead to a multiplicity of tortuous routes in three dimensions. 354

An external file that holds a picture, illustration, etc.
Object name is cr-2017-002594_0027.jpg

Fluctuations driving the autoionization of water involve the concerted motion of many atoms. A water wire links the H 3 O + and OH – ions. The Grotthuss mechanism (A) of proton propagation involves three consecutive steps (1–3). In contrast, a cooperative motion of the water wire (B) that results in a concerted motion of three protons is shown.

Water molecules can diffuse rapidly through nanotubes. It has been suggested that an excess of proton charge defects near the entrance of dry hydrophobic carbon nanotube can aid the loading of water. 355 Relying on charge defect delocalization and Grotthuss shuttling to help grow a chain of water molecules in the tube, this wetting process is specific to a hydrated excess proton alone. Other monatomic cations, such as K + , have the opposite effect: they block the wetting process and make the nanotube even more dry. A wetting mechanism of this type can be important in biological systems, for example in understanding the hydration of hydrophobic protein pores, where the charge defect is, for instance, created by peripheral amino acid residue deprotonation.

11. Challenges for Improving Water Models

To create new technologies for producing clean water, reclaiming polluted water, predicting climate and weather, inventing green chemistry, separating chemicals and biomolecules, and designing new drugs to cure diseases requires an ever deeper understanding of water structure–property relationships. It requires ever better models of various types and at different levels. Modeling is needed that can handle water that contains salts and oils and biomolecules, often in high concentrations, in the presence of complex and structured media and surfaces, and often under different conditions of temperature and pressure. Semiempirical models of water have generated insights and quantitative predictions over a broad spectrum of inquiries. But, there are opportunities for the future. More efficient computational models are needed to give better sampling of water’s distributions of configurations, including its cage-like states, to give more accurate entropies, heat capacities, and temperatures of transitions. We need improved water models for solvating ions of high charge density. We need continued work in polarizable models. We need to go beyond current fixed-charge models, if we are to study pH or acid–base behavior, because protons cannot dissociate in present models. We need more efficient quantum simulation models for acid–base chemistry, for bond-making and bond-breaking reactions, for water’s ices and solid states, and for understanding principles of bonding.

There is also great value in improved analytical and semianalytical and coarse-grained models. They can give insights into principles; they give ways to explore dependences on variables such as temperatures, pressures, and concentrations; and they should be able to give computationally efficient ways to address engineering questions in complex systems. Moreover, combining methods can also be valuable: quantum plus semiempirical modeling for bond-breaking reactions, or coarse-grained plus atomistic semiempirical models for noncovalent changes in large biomolecular complexes, for example.

12. Summary

We reviewed here how water’s properties are encoded in its molecular structure and energies, as interpreted through the lens of various models, theories, and computer simulations. Anomalous properties of water arise from the cage-like features of its molecular organization, arising from the tetrahedral hydrogen bonding among neighboring molecules. The challenge in modeling is due to the coupling between rotational and translational freedom of neighboring molecules. It is responsible for some volumetric anomalies, such as the lower density of ice than liquid water, the backward part of the liquid–solid pT phase diagram, and water’s aversion for nonpolar surfaces and solutes. Water’s behavior as a solvent for nonpolar and charged molecules can be explained through a combination of its caging structures and water’s electric dipole.

Acknowledgments

We appreciate the support of the Stony Brook University Laufer Center and of National Institutes of Health (NIH) Grant 5R01GM063592-15. M.L. and B.H.-L. acknowledge the financial support from the Slovenian Research Agency (ARRS) through research core funding No. P1-0201 and Project BI-US/16-17-045. M.L. acknowledges the partial founding of the ARRS Project L4-7628. M.F.-S. acknowledges funding from DOE Grant DE-FG02-09ER16052. We thank Thomas M. Truskett, Nico F. A. van der Vegt, Ben Hsiao, J. Paul Devlin, Judith Herzfeld, and Tomaž Urbič for helpful comments. We thank Sarina Bromberg for help with some figures.

Abbreviations

HBhydrogen bond
LJLennard-Jones
MOmolecular orbital
LUMOlowest unoccupied molecular orbital
HOMOhighest occupied molecular orbital
DFTdensity functional theory
HDLhigh density liquid
LDLlow density liquid
MBMercedes-Benz
TIPtransferable intermolecular potential
SPCsimple point charge
MDmolecular dynamics
PMFpotential of mean force
RDFradial distribution function
CIPcontact ion pair
SIPsolvent-shared ion pair
GBgeneralized Born
PBPoisson–Boltzmann
RISMreference interaction site model
SEAsemi-explicit assembly
i-PMFinterpolation potential of mean force
SAMPLStatistical Assessment of the Modeling of Proteins and Ligands

Biographies

Emiliano Brini received his M.Sc. degree in chemistry from the University of Bologna in 2008, and did a Ph.D. in chemistry with N. F. A. van der Vegt at Technische Universitaet Darmstadt. He is currently a postdoc with Ken A. Dill at the Laufer Center for Physical and Quantitative Biology at Stony Brook University. His research focuses on solvation and protein physics.

Christopher J. Fennell received his Ph.D. in chemistry from the University of Notre Dame, and after working as a postdoctoral researcher at the University of California, San Francisco, he became a Laufer Junior Fellow at the Laufer Center for Physical and Quantitative Biology at Stony Brook University. He is currently an assistant professor of chemistry at Oklahoma State University with a research focus in computational and theoretical chemistry.

Marivi Fernandez-Serra obtained her B.Sc. and M.Sc. in physics at Universidad Autonoma de Madrid, Spain, in 2001. She received her Ph.D. in physics in 2005, from the University of Cambridge. After this she worked as a postdoc at CECAM (Center for Atomic and Molecular Simulations) in Lyon, France. Currently she is an associate professor of physics and has been at Stony Brook University since 2008. Her research is in the field of computational condensed matter physics. She develops and applies methods to study the atomic and electronic dynamics of complex materials. One of her main research areas is the study of fundamental properties of liquid water using quantum mechanical simulations. In her group they apply their methods to study the interface between water and functional elements such as electrodes, photocatalytic semiconductors, and oxide materials.

Barbara Hribar-Lee studied chemistry at University of Ljubljana, Slovenia. She received her Ph.D. at the same university in 1998. In 2000–2001 she did her postdoctoral research with Ken A. Dill at the University of California, San Francisco. She is a professor of physical chemistry at the Faculty of Chemistry and Chemical Technology (FCCT), University of Ljubljana. Her research is in the fields of statistical mechanical theories of water and hydration, partly quenched systems, solutions of (poly)electrolytes, and intermolecular interactions and their role in the stability of protein solutions.

Miha Lukšič studied chemistry at the Faculty of Chemistry and Chemical Technology (FCCT), University of Ljubljana, Slovenia. In 2010 he obtained his Ph.D. in physical chemistry at the same university. He was a visiting researcher at the University of Regensburg (Germany), the National Autonomous University of Mexico (Mexico), and the University of California, San Francisco. In 2012–2014 he worked as a postdoctoral fellow at the Laufer Center for Physical and Quantitative Biology at Stony Brook University. He is currently an assistant professor of physical chemistry at the FCCT, University of Ljubljana. His research interests are in statistical mechanical theory of water and hydration, partly quenched systems, solutions of (poly)electrolytes, and lately in electrochromic cell devices.

Ken A. Dill received S.B. and S.M. degrees in mechanical engineering from the Massachusetts Institute of Technology, obtained a Ph.D. in biology with B. H. Zimm at the University of California, San Diego, and did postdoctoral research with P. J. Flory at Stanford University. He was on the faculty of the University of California, San Francisco, until 2010, and since then has been the Louis and Beatrice Laufer Professor of Physics and Chemistry at Stony Brook University and the Director of the Laufer Center for Physical and Quantitative Biology. His research has been on water statistical mechanics, the protein-folding problem, aspects of nonequilibrium statistical physics, and biophysical mechanisms and evolution of cells.

A.1. Modeling Large Complex Solutes Requires Approximations and Efficient Computational Methods

For studying water that contacts large or complex surfaces, efficient computational methods are needed. Several fast approximate models, described below, have been useful. 356 − 365

A.1.1. Modeling Water as a Continuum: Surface Tension, Born and Poisson Models

For modeling how water molecules interact with solutes, a simple and computationally inexpensive strategy is to treat water as a miniature unstructured continuous medium. For example, you can approximate the free energy of dissolving nonpolar solutes in water as Δ G = γ A , where A is the surface area of the solute and γ is the free energy per unit area of association of the solute with water (like a surface tension). In this approach, water’s role is modeled as a continuum represented by a single bulk-like quantity, γ in this case, rather than as individual molecules buffeted by thermal motions. Different values of γ are obtained for different types of chemical moieties by experiments on model compounds . For solutes having multiple chemical moieties, the total solvation free energy is assumed to be the sum of free energies of the component moieties. 185

A different continuum approximation is useful for treating the solvation of ions or polar molecules. For solutes that have electrostatic character, the relevant bulk-like property of water is its dielectric constant ε . When two charged objects interact with each other in a polarizable medium, such as liquid water, their interactions are weakened by the screening effect that results from the effect of the ions polarizing the medium. At room temperature, at charge–charge separations of greater than the Bjerrum length of water (roughly 7–8 Å), this screening is an effect of sheilding one charge from another. Implicit-solvent models treat water as a continuum, rather than as individual molecules. Implicit-solvent models treat water as having a dielectric constant, which reduces the ion–ion interaction energy. 356 − 363

For solute molecules that have both nonpolar and charged moieties, another type of additivity relationship is used. The free energy of solvation, Δ G solv , is taken to be the sum of electrostatic (Δ G el ) and nonpolar (Δ G nonpolar ) components. 364 , 365

A.1.1.1. Poisson Implicit-Solvent Models of Water

One type of implicit-solvent model is the Poisson or Poisson–Boltzmann approach. 366 − 370 In this case, all charge–charge interactions are taken into account through Coulomb’s law, summed together by solving the Poisson equation of electrostatics for the given distribution of fixed full and partial charges. In the presence of mobile charges , such as dissolved salt ions—which are free to distribute around the fixed charges—the Poisson–Boltzmann equation is solved for the electrostatic potential due to both fixed and mobile charges. 369 Poisson-type modeling requires choosing where to locate the dielectric boundaries, and what dielectric constants should be used to approximate the different regions. As implicit-solvent models go, solving the Poisson equation can be computationally expensive. 371 , 372

A.1.1.2. Born and Generalized Born Implicit-Solvent Models of Water

A less computationally expensive way to approximate the screening effect of water around charges is the generalized Born model. 357 , 364 , 373 , 374 In the underlying Born model, the electrostatic potential around a simple ion with radius a and charge q is 364

equation image

where ε w is the dielectric constant of water. In the case where a solute molecule can be modeled as several charges embedded in spheres, and the separation between them is sufficiently large compared to the radii, the solvation free energy of a molecule can be given by a sum of individual Born terms, and pairwise Coulomb terms. 364 In the generalized Born model the two terms are replaced by one, substituting the radius of the sphere and the distance between two charges by a single function, f GB , that interpolates between the two: 364

equation image

GB simulations tend to be faster than those of PB. To achieve similar accuracies, GB models are often parametrized on a corresponding PB model by choosing proper effective atomic radii. 375 − 378

The advantage of continuum models is their computational speed. Therefore, they are often applied in modeling large or complex solutes, such as biomolecules, or multicomponent systems, such as in chemical separations. 379 − 381 The continuum models sacrifice molecular detail or specific hydrogen bonding, 356 , 364 , 374 , 382 and thus water–water, water–solute, and solute–solute correlations. 383 For example, implicit models are often qualitatively incorrect in situations involving solvated charges near curved surfaces 268 or where solute surfaces are complex or corrugated. 384

A.1.2. Modeling Liquids Using the Reference Interaction Site Model (RISM)

A compromise between explicit-solvent and continuum models are the integral-equation theories of molecular liquids. They aim to combine a relatively detailed molecular description with relatively low computational cost. The fundamental relationship for the integral equation theories is the Ornstein–Zernike (OZ) integral equation which, for a m -component system, is 383

equation image

where r is the interparticle separation distance. Here, c ( r ) is the direct correlation function . The total correlation function is given by h ( r ) = g ( r ) – 1. To solve the OZ equation, another relation is needed between the functions c ( r ) and h ( r ), called a closure relation . Different closures are used for different problems. 383

One such approach to solvation in aqueous solutions is the reference interaction site model (RISM), proposed by Chandler and Andersen. 385 It treats the solute and solvent molecules as sets of sites having spherical symmetry, where the interactions between pairs of sites are treated through their correlation functions. RISM has been applied to dipolar liquids by Hirata and Rossky 386 (XRISM). The RISM theory has been used in combination with the hypernetted-chain (HNC) closure to study the solvation of monatomic solutes (alkali halides and argon-like particles) in aqueous solutions, 195 , 387 small peptides, 388 and polar organic molecules. 389 However, these early RISM models did not handle well large molecular solutes, which contain atoms that are buried inside away from solvent exposure, 390 or the polarizabilites of polar liquids.

Further developments led to the 3D-RISM model , in which a set of three-dimensional (3D) integral equations (one equation for each solvent site) was derived by integrating over the orientational degrees of freedom. 390 − 392 3D-RISM has been applied to studying the ionic atmosphere of DNA, 393 − 395 receptor–ligand pairs docking and binding affinities, 396 − 398 self-assembly of rosette nanotubes, 399 tubulin protofilaments, 400 and amyloid fibrils. 401 However, 3D-RISM gives limited accuracy in density distributions, 402 − 404 or hydration free energies of organic solutes, 405 or distribution coefficients for drug-like molecules between organic and aqueous phases. 406 These limitations have motivated more recent advances. 392 , 405 , 407 − 409 Misin et al. 408 demonstrated a pressure-corrected 3D-RISM that gives accurate salting-out constants for a wide range of organic compounds in aqueous sodium chloride solutions. Improved methods to predict hydration free energies and entropies of small drug-like molecules are now also available. 409 So far, the RISM approaches have been less widely used than implicit-solvent modeling because of the greater computational expense of RISM.

A.1.3. The SEA Water Model: Explicit-Model Physics at Implicit-Model Speeds

SEA (semi-explicit assembly) is a model that aims to compute solvation free energies using the physics of semiempirical models, such as TIP3P, but at much faster computational speeds. 186 , 410 − 412 In order to calculate the solvation free energy of a solute in water SEA uses two steps: (1) First is a pre-runtime simulation of idealized spheres (having different radii, different charges, and different van der Waals interactions with water) in explicit water, representing all the possible component pieces a solute could have. (2) Then, at runtime, SEA assembles the appropriate ideal spheres and sums terms into a total free energy of solvation. SEA is as fast as implicit-solvent models, because of its free-energy additivities in the assembly step. SEA has been found to be as accurate as explicit-solvent models in blind tests, called SAMPL; 413 − 415 see sec A.2 . While SEA is only as accurate as its underlying explicit-solvent model, nevertheless its accuracy stems, in part, from its intrinsic capture of water–water multibody effects in the presimulations. With the development of the field version of the technique and its dynamic solvent accessible surface boundary ( Figure ​ Figure28 28 ), SEA can also be applied to solvents beyond pure water by simply updating the surface accessibility and solvation coefficient tables for solvated LJ particles in the new environments. 412 , 415 , 416

An external file that holds a picture, illustration, etc.
Object name is cr-2017-002594_0028.jpg

The water-accessible surface depends on the charge on the solute (gray). (left) Water-accessible surface (green) around an uncharged solute. (right) When the solute is charged (blue), the water surface is electrostricted (red) around neighboring solute atoms.

A.1.4. The i-PMF Method for Computing Potentials of Mean Force

In the spirit of the SEA approach, i-PMF is a method for fast and accurate calculations of the potential of mean force (PMF) between pairs of charged or uncharged solutes in water. 417 The PMF is the reversible work averaged over all possible configurations of the solvent molecules required to bring two solute particles in the solution from infinite separation to a distance r . The i-PMF method is divided into presimulation and runtime steps. In the presimulation step, extensive MD simulations are performed with a classical force field to compute the PMFs of various pairs of solutes in the solvent. These computed PMFs are compiled into tables. Next, interpolations are performed to fill in additional solute radii and separations r . At runtime, a PMF can be computed rapidly for particles of arbitrary charges, interaction energies, and radii. Figure ​ Figure29 29 shows an example of Na + and Cl – ions in SPC/E water at infinite dilution. Figure ​ Figure29 29 also illustrates the nature and magnitude of the errors from the miniature-medium models, GB and PB, which do not account for the particulate nature of water, and from 3D-RISM, which treats waters in an explicit but simplified way. 418 Because i-PMF is an interpolator, its runtime speed is about 5 orders of magnitude faster than the underlying MD simulations that it interpolates.

An external file that holds a picture, illustration, etc.
Object name is cr-2017-002594_0029.jpg

Different models give different predictions for the Na + –Cl – PMF in water. 417 , 418

A.2. The SAMPL Competition for Modeling Solvation in Water

How can we evaluate and improve computational models of solvation? Different methods have different errors and different computational speeds. A community-wide mechanism has recently been developed for blind testing of solvation models. The event, called SAMPL (Statistical Assessment of the Modeling of Proteins and Ligands) is held every second year. In the SAMPL event, a set of small-molecule (often drug-like) structures is provided to the predictor community, who then uses their various methods to predict the solvation free energies. Experimental measurements are then performed subsequently, and the prior blind predictions of the different groups are then evaluated comparatively. 419 − 423

A.3. How Is Water Globally Distributed on Earth?

The total amount of water on Earth is around 1.39 × 10 9 km 3 ; see Table 2 . Of this, only around 0.77% is fresh water. Water makes up about 71% of the Earth’s surface, but only about 0.05% of the Earth’s total mass. Rivers are the source of most of the fresh surface water people use; that is about 1/10000th of 1% of Earth’s total water. Of Earth’s total water, 97.5% of it is saline, 95.5% of which is in oceans. Only 2.5% of all water is fresh water, and most of that is trapped as glaciers and snow.

water sourcewater volume [km ]fresh water [%]total water [%]
oceans, seas, bays1 338 000 00096.54
ice caps, glaciers, permanent snow24 064 00068.71.74
groundwater23 400 0001.69
fresh10 530 00030.10.76
saline12 870 0000.93
soil moisture16 5000.050.001
ground ice, permafrost300 0000.860.022
lakes176 4000.013
fresh91 0000.260.007
saline85 4000.006
atmosphere12 9000.040.001
swamp water11 4700.030.0008
rivers2 1200.0060.0002
biological water1 1200.0030.0001

A.4. Selected Physicochemical Properties of Liquid Water

Table 3 lists selected physicochemical properties of liquid water. The crystal structures and densities of various ice forms are listed in Table 4 .

propertyvalue
density997.047013 kg m
molar volume18.0685 cm  mol
molar concentration55.345 mol dm
molality55.508472 mol kg
dielectric constant78.375
magnetic susceptibility–1.64 × 10  m  mol
electric conductivity0.05501 × 10  Ω  cm
limiting molar conductivityH : 349.19 cm  Ω  mol
 OH : 199.24 cm  Ω  mol
ionic mobilityH : 3.623 × 10  m  V  s
 OH : 2.064 × 10  m  V  s
thermal conductivity0.610 W m  K
speed of sound1496.69922 m s
refractive index1.33286 (λ = 589.26 nm)
pH6.9976
p 13.995
surface tension0.07198 N m
kinematic viscosity0.8935 × 10  m  s
dynamic viscosity0.8909 mPa s
bulk viscosity2.47 mPa s
diffusion coefficient0.2299 Å  ps
dipole moment2.95 D (at 27 °C)
adiabatic compressibility0.4477 GPa
isothermal compressibility0.4599 GPa
expansion coefficient0.000253 °C
adiabatic elasticity2.44 GPa
Joule–Thomson coefficient0.214 K MPa
vapor pressure3.165 kPa
cryoscopic constant1.8597 K kg mol
ebullioscopic constant0.5129 K kg mol
polarizability1.636 × 10  F m
ice formcrystal structuredensity [g cm ]
I hexagonal0.92
I cubic0.93
IIrhombohedral1.17
IIItetragonal1.14
IVrhombohedral1.27
Vmonoclinic1.23
VItetragonal1.31
VIIcubic1.50
VIIItetragonal1.46
IXtetragonal1.16
Xcubic2.51
XIorthorhombic0.92
XIItetragonal1.29
XIIImonoclinic1.23
XIVorthorhombic1.29
XVpseudoorthorhombic1.30
XVIcubic0.81

A.5. Parameters for Some Water Models

The parameters for some water molecular models are listed in Table 5 . Different water models used in Table 5 are shown in Figure ​ Figure30 30 .

An external file that holds a picture, illustration, etc.
Object name is cr-2017-002594_0030.jpg

Different water models used in Table 5 .

modeltypeσ [Å]ε [kJ mol ]  [Å]  [Å]  [ ]  [ ]θ [deg]ϕ [deg]
SPCa3.1660.6510.41–0.82109.47
SPC/Ea3.1660.6510.4238–0.8476109.47
TIP3Pa3.150610.63640.95720.417–0.834104.52
iAMOEBA a3.64530.82350.95840.29701–0.59402106.48
TIP4Pb3.153650.6480.95720.150.52–1.04104.5252.26
TIP4P-Ewb3.164350.6809460.95720.1250.52422–1.04844104.5252.26
TIP4P/2005b3.15890.77490.95720.15460.5564–1.1128104.5252.26
ST2c3.10.3169410.80.24357–0.24357109.47109.47
TIP5Pc3.120.66940.95720.70.241–0.241104.52109.47
TIP5P-Ewc3.0970.74480.95720.70.241–0.241104.52109.47
modeldipole moment [D]dielectric constself-diffusion, 10  [cm  s ]density max [°C]expansion coeff, 10  [°C ]
SPC2.27653.85–457.3
SPC/E2.35712.49–385.14
TIP3P2.35825.19–919.2
iAMOEBA2.7880.72.5442.5
TIP4P2.18533.29–254.4
TIP4P-Ew2.3262.92.413.1
TIP4P/20052.305602.0852.8
TIP5P2.2981.52.6246.3
TIP5P-Ew2.29922.884.9
exptl2.9578.42.33.9842.53

The water–water pair interaction potential ( u ww ) for non-polarizable pointcharge water models is calculated using the parameters from Table 5 through eq 4 :

equation image

where r OO is the distance between two oxygens, and r ij is a distance between charges q i and q j , located on different water molecules.

A.6. Calculated Physicochemical Properties of Some Water Models

The calculated physicochemical properties of some water models are listed in Table 6 .

The authors declare no competing financial interest.

Author Status

# C.J.F., M.F.-S., B.H.-L., and M.L.: Authors are listed in alphabetical order.

  • Coonfield T. The Varmits: Living With Appalachian Outlaws ; AuthorHouse: Bloomington, IN, 2011; p 48. [ Google Scholar ]
  • USGS—U.S. Geological Survey. How Much Water Is There On, In, and Above the Earth? 2016. http://water.usgs.gov/edu/earthhowmuch.html (accessed September 15, 2016).
  • Shiklomanov I. In Water in Crisis: A Guide to the World’s Fresh Water Resources ; Gleick P. H., Ed.; Oxford University Press: New York, 1993; pp 13–25. [ Google Scholar ]
  • Brown P. S.; Bhushan B. Bioinspired Materials for Water Supply and Management: Water Collection, Water Purification and Separation of Water From Oil . Philos. Trans. R. Soc., A 2016, 374 , 20160135. 10.1098/rsta.2016.0135. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Caldecott J. Water: Life in Every Drop ; Virgin Digital: 2008. [ Google Scholar ]
  • Henry M. The State of Water in Living Systems: From the Liquid to the Jellyfish . Cell. Mol. Biol. 2005, 51 , 677–702. [ PubMed ] [ Google Scholar ]
  • Ling G. What Determines the Normal Water Content of a Living Cell? . Physiol. Chem. Phys. Med. NMR 2004, 36 , 1–19. [ PubMed ] [ Google Scholar ]
  • Tree Physics 1: Capillary Action, the Height of Trees, and the Optimal Placement of Branches; 2016. https://npand.wordpress.com/2008/08/05/tree-physics-1/ (accessed September 15, 2016).
  • United Nations, Department of Economic and Social Affairs, Population Division. Population 2030: Demographic Challenges and Opportunities for Sustainable Development Planning ; ST/ESA/SER.A/389; United Nations: 2015.
  • Water Resources Group. Charting Our Water Future: Economic Frameworks to Inform Decision-Making; 2009. http://www.2030wrg.org/wp-content/uploads/2014/07/Charting-Our-Water-Future-Final.pdf (accessed September 15, 2016).
  • United Nations. Water; 2016. http://www.un.org/en/sections/issues-depth/water/index.html (accessed September 15, 2016).
  • OECD. Environmental Outlook to 2030 . OECD Publishing: Paris, France, 2008. [ Google Scholar ]
  • Greenlee L. F.; Lawler D. F.; Freeman B. D.; Marrot B.; Moulin P. Reverse Osmosis Desalination: Water Sources, Technology, and Today’s Challenges . Water Res. 2009, 43 , 2317–2348. 10.1016/j.watres.2009.03.010. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Subramani A.; Jacangelo J. G. Emerging Desalination Technologies for Water Treatment: A Critical Review . Water Res. 2015, 75 , 164–187. 10.1016/j.watres.2015.02.032. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Gleick P. Water and Conflict . Int. Security 1993, 18 , 79–112. 10.2307/2539033. [ CrossRef ] [ Google Scholar ]
  • Pacific Institute. Water Conflict Chronology List; 2016. http://worldwater.org/water-conflict (accessed September 15, 2016).
  • Masahiro M. Managing Water for Peace in the Middle East: Alternative Strategies ; United Nations University Press: New York, 1995. [ Google Scholar ]
  • NASA. Global Water Budget; 2016. http://neptune.gsfc.nasa.gov/index.php?section=26 .
  • Seinfeld J. H.; Pandis S. N.. Atmospheric Chemistry and Physics: From Air Pollution to Climate Change . (3rd Ed.), J. Wiley & Sons: Hoboken, 2016. .
  • Infoplease. Water Pollution; 2016. http://www.infoplease.com/encyclopedia/science/water-pollution-industrial-pollution.html (accessed September 15, 2016).
  • Petrie B.; Barden R.; Kasprzyk-Hordern B. A Review on Emerging Contaminants in Wastewaters and the Environment: Current Knowledge, Understudied Areas and Recommendations for Future Monitoring . Water Res. 2015, 72 , 3–27. 10.1016/j.watres.2014.08.053. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Qu X.; Alvarez P. J.; Li Q. Applications of Nanotechnology in Water and Wastewater Treatment . Water Res. 2013, 47 , 3931–3946. 10.1016/j.watres.2012.09.058. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Ahuja S.; Hristovski K. In Novel Solutions to Water Pollution ; Satinder A., Kiril H., Eds.; American Chemical Society: Washington, DC, 2013. [ Google Scholar ]
  • Breslow R. Handbook of Green Chemistry ; Wiley-VCH Verlag GmbH & Co. KGaA: 2010. [ Google Scholar ]
  • Horváth I. T.; Anastas P. T. Innovations and Green Chemistry . Chem. Rev. 2007, 107 , 2169–2173. 10.1021/cr078380v. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Clean Technica. New Water Treatment Process Enhances Oil Recovery; 2016. https://cleantechnica.com/2013/03/05/fracking-water-new-water-treatment-process-enhances-oil-recovery-removes-99-of-org (accessed September 15, 2016).
  • Kuchment A. Drilling for Earthquakes . Sci. Am. 2016, 315 , 46–53. 10.1038/scientificamerican0716-46. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Wagner W.; Pruß A. The IAPWS Formulation 1995 for the Thermodynamic Properties of Ordinary Water Substance for General and Scientific Use . J. Phys. Chem. Ref. Data 2002, 31 , 387–535. 10.1063/1.1461829. [ CrossRef ] [ Google Scholar ]
  • Ben-Naim A. Statistical Thermodynamics for Chemists and Biochemists ; Springer: 1992; pp 459–559. [ Google Scholar ]
  • Eisenberg D.; Kauzmann W.. The Structure and Properties of Water ; Oxford University Press on Demand: 2005. [ Google Scholar ]
  • Franks F. Water: A Comprehensive Treatise. Volume 1. The Physics and Physical Chemistry of Water ; Plenum Press: 1972. [ Google Scholar ]
  • Franks F. Water: A Matrix of Life ; Royal Society of Chemistry: 2000; Vol. 21 . [ Google Scholar ]
  • Franks F. The Physics and Physical Chemistry of Water ; Springer Science & Business Media: 2012; Vol. 1 . [ Google Scholar ]
  • Guillot B. A Reappraisal of What We Have Learnt During Three Decades of Computer Simulations on Water . J. Mol. Liq. 2002, 101 , 219–260. 10.1016/S0167-7322(02)00094-6. [ CrossRef ] [ Google Scholar ]
  • Ben-Naim A. Molecular Theory of Water and Aqueous Solutions. Part I: Understanding Water ; World Scientific: Singapore, 2009. [ Google Scholar ]
  • Ben-Naim A. Molecular Theory of Water and Aqueous Solutions: Part II: The Role of Water in Protein Folding, Self-Assembly and Molecular Recognition ; World Scientific: Singapore, 2011. [ Google Scholar ]
  • Chaplin M. Water Structure and Science; 2016. http://www1.lsbu.ac.uk/water/ (accessed September 15, 2016).
  • Sun C. Q.; Sun Y. In Springer Series in Chemical Physics ; Castleman A. W., Toennies J. P., Yamanouchi K., Zinth W., Eds.; Springer: Singapore, 2016; Vol. 113 . [ Google Scholar ]
  • Dill K. A., Bromberg S.. Molecular Driving Forces: Statistical Thermodynamics in Biology, Chemistry, Physics, and Nanoscience , 2nd ed.; Garland Science: New York, 2010. [ Google Scholar ]
  • Dill K. A.; Truskett T. M.; Vlachy V.; Hribar-Lee B. Modeling Water, the Hydrophobic Effect, and Ion Solvation . Annu. Rev. Biophys. Biomol. Struct. 2005, 34 , 173–199. 10.1146/annurev.biophys.34.040204.144517. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Röntgen W. C. Ueber die Constitution des flüssigen Wassers . Ann. Phys. 1892, 281 , 91–97. 10.1002/andp.18922810108. [ CrossRef ] [ Google Scholar ]
  • Bernal J. D.; Fowler R. H. A Theory of Water and Ionic Solution, With Particular Reference to Hydrogen and Hydroxyl Ions . J. Chem. Phys. 1933, 1 , 515–548. 10.1063/1.1749327. [ CrossRef ] [ Google Scholar ]
  • Pauling L. The Nature of the Chemical Bond , 2nd ed.; Cornell University: Ithaca, NY, 1939,. [ Google Scholar ]
  • Samoilov O. I. Struktura Vodnykh Rastvorov Elektrolitov I Gidratatsiia Ionov ; Izd-vo Akademii nauk SSSR: Moscow, 1957. [ Google Scholar ]
  • Pople J. A. Molecular Association in Liquids. II. A Theory of the Structure of Water . Proc. R. Soc. London, Ser. A 1951, 205 , 163–178. 10.1098/rspa.1951.0024. [ CrossRef ] [ Google Scholar ]
  • Romero A. H.; Silvestrelli P. L.; Parrinello M. Compton Scattering and the Character of the Hydrogen Bond in Ice I h . J. Chem. Phys. 2001, 115 , 115–123. 10.1063/1.1372183. [ CrossRef ] [ Google Scholar ]
  • Burnham C. J.; Li J.; Xantheas S. S.; Leslie M. The Parametrization of a Thole-Type All-Atom Polarizable Water Model From First Principles and Its Application to the Study of Water Clusters (N = 2–21) and the Phonon Spectrum of Ice Ih . J. Chem. Phys. 1999, 110 , 4566–4581. 10.1063/1.478797. [ CrossRef ] [ Google Scholar ]
  • Sprik M.; Klein M. L. A Polarizable Model for Water Using Distributed Charge Sites . J. Chem. Phys. 1988, 89 , 7556–7560. 10.1063/1.455722. [ CrossRef ] [ Google Scholar ]
  • Zhu S. B.; Yao S.; Zhu J. B.; Singh S.; Robinson G. W. A Flexible/Polarizable Simple Point Charge Water Model . J. Phys. Chem. 1991, 95 , 6211–6217. 10.1021/j100169a031. [ CrossRef ] [ Google Scholar ]
  • Lamoureux G.; MacKerell A. D. Jr.; Roux B. A Simple Polarizable Model of Water Based on Classical Drude Oscillators . J. Chem. Phys. 2003, 119 , 5185–5197. 10.1063/1.1598191. [ CrossRef ] [ Google Scholar ]
  • Yu H.; Hansson T.; van Gunsteren W. F. Development of a Simple, Self-Consistent Polarizable Model for Liquid Water . J. Chem. Phys. 2003, 118 , 221–234. 10.1063/1.1523915. [ CrossRef ] [ Google Scholar ]
  • Kumar R.; Wang F.-F.; Jenness G. R.; Jordan K. D. A Second Generation Distributed Point Polarizable Water Model . J. Chem. Phys. 2010, 132 , 014309. 10.1063/1.3276460. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Wang L.-P.; Head-Gordon T.; Ponder J. W.; Ren P.; Chodera J. D.; Eastman P. K.; Martinez T. J.; Pande V. S. Systematic Improvement of a Classical Molecular Model of Water . J. Phys. Chem. B 2013, 117 , 9956–9972. 10.1021/jp403802c. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Tröster P.; Lorenzen K.; Schwörer M.; Tavan P. Polarizable Water Models From Mixed Computational and Empirical Optimization . J. Phys. Chem. B 2013, 117 , 9486–9500. 10.1021/jp404548k. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Xantheas S. S. Cooperativity and Hydrogen Bonding Network in Water Clusters . Chem. Phys. 2000, 258 , 225–231. 10.1016/S0301-0104(00)00189-0. [ CrossRef ] [ Google Scholar ]
  • Elton D. C.; Fernández-Serra M.-V. Polar Nanoregions in Water: A Study of the Dielectric Properties of TIP4P/2005, TIP4P/2005f and TTM3F . J. Chem. Phys. 2014, 140 , 124504. 10.1063/1.4869110. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Hasegawa T.; Tanimura Y. A Polarizable Water Model for Intramolecular and Intermolecular Vibrational Spectroscopies . J. Phys. Chem. B 2011, 115 , 5545–5553. 10.1021/jp111308f. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Pamuk B.; Soler J. M.; Ramírez R.; Herrero C. P.; Stephens P. W.; Allen P. B.; Fernández-Serra M.-V. Anomalous Nuclear Quantum Effects in Ice . Phys. Rev. Lett. 2012, 108 , 193003. 10.1103/PhysRevLett.108.193003. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Mahoney M. W.; Jorgensen W. L. Quantum, Intramolecular Flexibility, and Polarizability Effects on the Reproduction of the Density Anomaly of Liquid Water by Simple Potential Functions . J. Chem. Phys. 2001, 115 , 10758–10768. 10.1063/1.1418243. [ CrossRef ] [ Google Scholar ]
  • Yu H.; van Gunsteren W. F. Charge-On-Spring Polarizable Water Models Revisited: From Water Clusters to Liquid Water to Ice . J. Chem. Phys. 2004, 121 , 9549–9564. 10.1063/1.1805516. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Xu H.; Stern H. A.; Berne B. J. Can Water Polarizability Be Ignored in Hydrogen Bond Kinetics? . J. Phys. Chem. B 2002, 106 , 2054–2060. 10.1021/jp013426o. [ CrossRef ] [ Google Scholar ]
  • Car R.; Parrinello M. Unified Approach for Molecular Dynamics and Density-Functional Theory . Phys. Rev. Lett. 1985, 55 , 2471–2474. 10.1103/PhysRevLett.55.2471. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Laasonen K.; Sprik M.; Parrinello M.; Car R. ”Ab Initio” Liquid Water . J. Chem. Phys. 1993, 99 , 9080–9089. 10.1063/1.465574. [ CrossRef ] [ Google Scholar ]
  • Barker J. A.; Watts R. O. Structure of Water; A Monte Carlo Calculation . Chem. Phys. Lett. 1969, 3 , 144–145. 10.1016/0009-2614(69)80119-3. [ CrossRef ] [ Google Scholar ]
  • Rowlinson J. S. The Lattice Energy of Ice and the Second Virial Coefficient of Water Vapour . Trans. Faraday Soc. 1951, 47 , 120–129. 10.1039/tf9514700120. [ CrossRef ] [ Google Scholar ]
  • Rahman A.; Stillinger F. H. Molecular Dynamics Study of Liquid Water . J. Chem. Phys. 1971, 55 , 3336–3359. 10.1063/1.1676585. [ CrossRef ] [ Google Scholar ]
  • Ben-Naim A.; Stillinger F. H.. Aspects of the Statistical-Mechanical Theory of Water . In Structure and Transport Processes in Water and Aqueous Solutions ; Horne R. A., Ed.; Wiley-Interscience: New York, 1972. [ Google Scholar ]
  • Stillinger F. H.; Rahman A. Improved Simulation of Liquid Water by Molecular Dynamics . J. Chem. Phys. 1974, 60 , 1545–1557. 10.1063/1.1681229. [ CrossRef ] [ Google Scholar ]
  • Stillinger F. H. Theory and Molecular Models for Water . Adv. Chem. Phys. 1975, 31 , 1–101. 10.1002/9780470143834.ch1. [ CrossRef ] [ Google Scholar ]
  • Stillinger F. H.; Weber T. A. Inherent Structure in Water . J. Phys. Chem. 1983, 87 , 2833–2840. 10.1021/j100238a027. [ CrossRef ] [ Google Scholar ]
  • Fernández-Serra M. V.; Artacho E. Electrons and Hydrogen-Bond Connectivity in Liquid Water . Phys. Rev. Lett. 2006, 96 , 016404. 10.1103/PhysRevLett.96.016404. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Jorgensen W. L. Quantum and Statistical Mechanical Studies of Liquids. 10. Transferable Intermolecular Potential Functions for Water, Alcohols, and Ethers. Application to Liquid Water . J. Am. Chem. Soc. 1981, 103 , 335–340. 10.1021/ja00392a016. [ CrossRef ] [ Google Scholar ]
  • Jorgensen W. L.; Chandrasekhar J.; Madura J. D.; Impey R. W.; Klein M. L. Comparison of Simple Potential Functions for Simulating Liquid Water . J. Chem. Phys. 1983, 79 , 926–935. 10.1063/1.445869. [ CrossRef ] [ Google Scholar ]
  • Berendsen H. J. C.; Grigera J. R.; Straatsma T. P. The Missing Term in Effective Pair Potentials . J. Phys. Chem. 1987, 91 , 6269–6271. 10.1021/j100308a038. [ CrossRef ] [ Google Scholar ]
  • Berendsen H. J. C.; Postma J. P. M.; van Gunsteren W. F.; Hermans J. In Intermolecular Forces ; Pullman B., Ed.; Reidel: Dordrecht, 1981; pp 331–342. [ Google Scholar ]
  • Scott W. R. P.; Hünenberger P. H.; Tironi I. G.; Mark A. E.; Billeter S. R.; Fennen J.; Torda A. E.; Huber T.; Krüger P.; van Gunsteren W. F. The GROMOS Biomolecular Simulation Program Package . J. Phys. Chem. A 1999, 103 , 3596–3607. 10.1021/jp984217f. [ CrossRef ] [ Google Scholar ]
  • Brooks B. R.; Bruccoleri R. E.; Olafson B. D.; States D. J.; Swaminathan S.; Karplus M. CHARMM: A Program for Macromolecular Energy, Minimization, and Dynamics Calculations . J. Comput. Chem. 1983, 4 , 187–217. 10.1002/jcc.540040211. [ CrossRef ] [ Google Scholar ]
  • Pearlman D. A.; Case D. A.; Caldwell J. W.; Ross W. S.; Cheatham T. E. III; DeBolt S.; Ferguson D.; Seibel G.; Kollman P. AMBER a Package of Computer Programs for Applying Molecular Mechanics, Normal Mode Analysis, Molecular Dynamics, and Free Energy Calculations to Simulate the Structural and Energetic Properties of Molecules . Comput. Phys. Commun. 1995, 91 , 1–41. 10.1016/0010-4655(95)00041-D. [ CrossRef ] [ Google Scholar ]
  • Horn H. W.; Swope W. C.; Pitera J. W.; Madura J. D.; Dick T. J.; Hura G. L.; Head-Gordon T. Development of an Improved Four-Site Water Model for Biomolecular Simulations: TIP4P-Ew . J. Chem. Phys. 2004, 120 , 9665–9678. 10.1063/1.1683075. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Abascal J. L. F.; Vega C. A General Purpose Model for the Condensed Phases of Water: TIP4P/2005 . J. Chem. Phys. 2005, 123 , 234505. 10.1063/1.2121687. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Wang L.-P.; Martinez T. J.; Pande V. S. Building Force Fields: An Automatic, Systematic, and Reproducible Approach . J. Phys. Chem. Lett. 2014, 5 , 1885–1891. 10.1021/jz500737m. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Molinero V.; Moore E. B. Water Modeled as an Intermediate Element Between Carbon and Silicon . J. Phys. Chem. B 2009, 113 , 4008–4016. 10.1021/jp805227c. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Dhabal D.; Chakravarty C.; Molinero V.; Kashyap H. K. Comparison of Liquid-State Anomalies in Stillinger-Weber Models of Water, Silicon, and Germanium . J. Chem. Phys. 2016, 145 , 214502. 10.1063/1.4967939. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Espinosa J. R.; Navarro C.; Sanz E.; Valeriani C.; Vega C. On the Time Required to Freeze Water . J. Chem. Phys. 2016, 145 , 211922. 10.1063/1.4965427. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Xu L.; Molinero V. Is There a Liquid-Liquid Transition in Confined Water? . J. Phys. Chem. B 2011, 115 , 14210–14216. 10.1021/jp205045k. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Song B.; Molinero V. Thermodynamic and Structural Signatures of Water-Driven Methane-Methane Attraction in Coarse-Grained mW Water . J. Chem. Phys. 2013, 139 , 054511. 10.1063/1.4816005. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Nguyen A. H.; Molinero V. Identification of Clathrate Hydrates, Hexagonal Ice, Cubic Ice, and Liquid Water in Simulations: The CHILL+ Algorithm . J. Phys. Chem. B 2015, 119 , 9369–9376. 10.1021/jp510289t. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Liu Y.; Ichiye T. Soft Sticky Dipole Potential for Liquid Water: A New Model . J. Phys. Chem. 1996, 100 , 2723–2730. 10.1021/jp952324t. [ CrossRef ] [ Google Scholar ]
  • Bratko D.; Blum L.; Luzar A. A simple model for the intermolecular potential of water . J. Chem. Phys. 1985, 83 , 6367–6370. 10.1063/1.449585. [ CrossRef ] [ Google Scholar ]
  • Chandra A.; Ichiye T. Dynamical properties of the soft sticky dipole model of water: Molecular dynamics simulations . J. Chem. Phys. 1999, 111 , 2701–2709. 10.1063/1.479546. [ CrossRef ] [ Google Scholar ]
  • Tan M.-L.; Fischer J. T.; Chandra A.; Brooks B. R.; Ichiye T. A temperature of maximum density in soft sticky dipole water . Chem. Phys. Lett. 2003, 376 , 646–652. 10.1016/S0009-2614(03)01044-3. [ CrossRef ] [ Google Scholar ]
  • Fennell C. J.; Gezelter J. D. On the structural and transport properties of the soft sticky dipole and related single point water models . J. Chem. Phys. 2004, 120 , 9175–9184. 10.1063/1.1697381. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Ichiye T.; Tan M.-L. Soft Sticky Dipole-Quadrupole-Octupole Potential Energy Function for Liquid Water: An Approximate Moment Expansion . J. Chem. Phys. 2006, 124 , 134504. 10.1063/1.2161201. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Te J. A.; Ichiye T. Temperature and Pressure Dependence of the Optimized Soft-Sticky Dipole-Quadrupole-Octupole Water Model . J. Chem. Phys. 2010, 132 , 114511. 10.1063/1.3359432. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Fennell C. J.; Gezelter J. D. Computational Free Energy Studies of a New Ice Polymorph Which Exhibits Greater Stability Than Ice I h . J. Chem. Theory Comput. 2005, 1 , 662–667. 10.1021/ct050005s. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Senapati S.; Chandra A. Dielectric Constant of Water Confined in a Nanocavity . J. Phys. Chem. B 2001, 105 , 5106–5109. 10.1021/jp011058i. [ CrossRef ] [ Google Scholar ]
  • Tan M.-L.; Lucan L.; Ichiye T. Study of Multipole Contributions to the Structure of Water Around Ions in Solution Using the Soft Sticky Dipole-Quadrupole-Octupole (SSDQO) Model of Water . J. Chem. Phys. 2006, 124 , 174505. 10.1063/1.2177240. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Rodgers J. M.; Ichiye T. Multipole moments of water molecules and the aqueous solvation of monovalent ions . J. Mol. Liq. 2017, 228 , 54–62. 10.1016/j.molliq.2016.10.007. [ CrossRef ] [ Google Scholar ]
  • Izvekov S.; Voth G. A. Multiscale Coarse Graining of Liquid-State Systems . J. Chem. Phys. 2005, 123 , 134105. 10.1063/1.2038787. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Izvekov S.; Voth G. A. Multiscale Coarse-Graining of Mixed Phospholipid/Cholesterol Bilayers . J. Chem. Theory Comput. 2006, 2 , 637–648. 10.1021/ct050300c. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Buldyrev S. V.; Kumar P.; Debenedetti P. G.; Rossky P. J.; Stanley H. E. Water-Like Solvation Thermodynamics in a Spherically Symmetric Solvent Model With Two Characteristic Lengths . Proc. Natl. Acad. Sci. U. S. A. 2007, 104 , 20177–20182. 10.1073/pnas.0708427104. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Shell M. S. The Relative Entropy Is Fundamental to Multiscale and Inverse Thermodynamic Problems . J. Chem. Phys. 2008, 129 , 144108–144108. 10.1063/1.2992060. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Wang H.; Junghans C.; Kremer K. Comparative Atomistic and Coarse-Grained Study of Water: What do We Lose by Coarse-Graining? . Eur. Phys. J. E: Soft Matter Biol. Phys. 2009, 28 , 221–229. 10.1140/epje/i2008-10413-5. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Lyubartsev A.; Mirzoev A.; Chen L.; Laaksonen A. Systematic Coarse-Graining of Molecular Models by the Newton Inversion Method . Faraday Discuss. 2010, 144 , 43–56. 10.1039/B901511F. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Larini L.; Lu L.; Voth G. A. The Multiscale Coarse-Graining Method. VI. Implementation of Three-Body Coarse-Grained Potentials . J. Chem. Phys. 2010, 132 , 164107. 10.1063/1.3394863. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Brini E.; Algaer E. A.; Ganguly P.; Li C.; Rodríguez-Ropero F.; van der Vegt N. F. Systematic Coarse-Graining Methods for Soft Matter Simulations-A Review . Soft Matter 2013, 9 , 2108–2119. 10.1039/C2SM27201F. [ CrossRef ] [ Google Scholar ]
  • Hoogerbrugge P.; Koelman J. Simulating Microscopic Hydrodynamic Phenomena With Dissipative Particle Dynamics . Europhys. Lett. 1992, 19 , 155. 10.1209/0295-5075/19/3/001. [ CrossRef ] [ Google Scholar ]
  • Espanol P.; Warren P. Statistical Mechanics of Dissipative Particle Dynamics . Europhys. Lett. 1995, 30 , 191. 10.1209/0295-5075/30/4/001. [ CrossRef ] [ Google Scholar ]
  • Marrink S. J.; Risselada H. J.; Yefimov S.; Tieleman D. P.; De Vries A. H. The MARTINI Force Field: Coarse Grained Model for Biomolecular Simulations . J. Phys. Chem. B 2007, 111 , 7812–7824. 10.1021/jp071097f. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Yesylevskyy S. O.; Schäfer L. V.; Sengupta D.; Marrink S. J. Polarizable Water Model for the Coarse-Grained MARTINI Force Field . PLoS Comput. Biol. 2010, 6 , e1000810. 10.1371/journal.pcbi.1000810. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Jagla E. A. The Interpretation of Water Anomalies in Terms of Core-Softened Models . Braz. J. Phys. 2004, 34 , 17–23. 10.1590/S0103-97332004000100003. [ CrossRef ] [ Google Scholar ]
  • Chaimovich A.; Shell M. S. Anomalous Waterlike Behavior in Spherically-Symmetric Water Models Optimized With the Relative Entropy . Phys. Chem. Chem. Phys. 2009, 11 , 1901–1915. 10.1039/b818512c. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Vilaseca P.; Franzese G. Isotropic Soft-Core Potentials With Two Characteristic Length Scales and Anomalous Behaviour . J. Non-Cryst. Solids 2011, 357 , 419–426. 10.1016/j.jnoncrysol.2010.07.053. [ CrossRef ] [ Google Scholar ]
  • Torres-Carbajal A.; Castaneda-Priego R. Characterisation of the Thermodynamics, Structure and Dynamics of a Water-Like Model in 2- And 3-Dimensions . Phys. Chem. Chem. Phys. 2016, 18 , 17335–17340. 10.1039/C6CP01565D. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Kumar R.; Skinner J. L. Water Simulation Model with Explicit Three-Molecule Interactions . J. Phys. Chem. B 2008, 112 , 8311–8318. 10.1021/jp8009468. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Babin V.; Leforestier C.; Paesani F. Development of a “First Principles” Water Potential With Flexible Monomers: Dimer Potential Energy Surface, VRT Spectrum, and Second Virial Coefficient . J. Chem. Theory Comput. 2013, 9 , 5395–5403. 10.1021/ct400863t. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Babin V.; Medders G. R.; Paesani F. Development of a “First Principles” Water Potential With Flexible Monomers. II: Trimer Potential Energy Surface, Third Virial Coefficient, and Small Clusters . J. Chem. Theory Comput. 2014, 10 , 1599–1607. 10.1021/ct500079y. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Fennell C. J.; Li L.; Dill K. A. Simple Liquid Models With Corrected Dielectric Constants . J. Phys. Chem. B 2012, 116 , 6936–6944. 10.1021/jp3002383. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Izadi S.; Onufriev A. V. Accuracy Limit of Rigid 3-Point Water Models . J. Chem. Phys. 2016, 145 , 074501. 10.1063/1.4960175. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Cisneros G. A.; Wikfeldt K. T.; Ojamäe L.; Lu J.; Xu Y.; Torabifard H.; Bartók A. P.; Csányi G.; Molinero V.; Paesani F. Modeling Molecular Interactions in Water: From Pairwise to Many-Body Potential Energy Functions . Chem. Rev. 2016, 116 , 7501–7528. 10.1021/acs.chemrev.5b00644. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Ben-Naim A. Statistical Mechanics of “Waterlike” Particles in Two Dimensions. I. Physical Model and Application of the Percus-Yevick Equation . J. Chem. Phys. 1971, 54 , 3682–3695. 10.1063/1.1675414. [ CrossRef ] [ Google Scholar ]
  • Silverstein K. A. T.; Haymet A. D. J.; Dill K. A. A Simple Model of Water and the Hydrophobic Effect . J. Am. Chem. Soc. 1998, 120 , 3166–3175. 10.1021/ja973029k. [ CrossRef ] [ Google Scholar ]
  • Silverstein K. A. T.; Haymet A. D. J.; Dill K. A. Molecular Model of Hydrophobic Solvation . J. Chem. Phys. 1999, 111 , 8000–8009. 10.1063/1.480133. [ CrossRef ] [ Google Scholar ]
  • Silverstein K. A. T.; Dill K. A.; Haymet A. D. J. Hydrophobicity in a Simple Model of Water: Entropy Penalty as a Sum of Competing Terms via Full, Angular Expansion . J. Chem. Phys. 2001, 114 , 6303–6314. 10.1063/1.1355997. [ CrossRef ] [ Google Scholar ]
  • Hribar B.; Southall N. T.; Vlachy V.; Dill K. A. How Ions Affect the Structure of Water . J. Am. Chem. Soc. 2002, 124 , 12302–12311. 10.1021/ja026014h. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Urbič T.; Vlachy V.; Kalyuzhnyi Y. V.; Southall N. T.; Dill K. A. A Two-Dimensional Model of Water: Theory and Computer Simulations . J. Chem. Phys. 2000, 112 , 2843–2848. 10.1063/1.480928. [ CrossRef ] [ Google Scholar ]
  • Urbič T.; Vlachy V.; Kalyuzhnyi Y. V.; Dill K. A. Orientation-Dependent Integral Equation Theory for a Two-Dimensional Model of Water . J. Chem. Phys. 2003, 118 , 5516–5525. 10.1063/1.1556754. [ CrossRef ] [ Google Scholar ]
  • Truskett T. M.; Dill K. A. A Simple Statistical Mechanical Model of Water . J. Phys. Chem. B 2002, 106 , 11829–11842. 10.1021/jp021418h. [ CrossRef ] [ Google Scholar ]
  • Truskett T. M.; Dill K. A. Predicting Water’s Phase Diagram and Liquid-State Anomalies . J. Chem. Phys. 2002, 117 , 5101–5104. 10.1063/1.1505438. [ CrossRef ] [ Google Scholar ]
  • Urbic T.; Dill K. A. A Statistical Mechanical Theory for a Two-Dimensional Model of Water . J. Chem. Phys. 2010, 132 , 224507. 10.1063/1.3454193. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Urbic T. Analytical Model for Three-Dimensional Mercedes-Benz Water Molecules . Phys. Rev. E 2012, 85 , 061503. 10.1103/PhysRevE.85.061503. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Lukşič M.; Urbic T.; Hribar-Lee B.; Dill K. A. Simple Model of Hydrophobic Hydration . J. Phys. Chem. B 2012, 116 , 6177–6186. 10.1021/jp300743a. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Williamson C. H.; Hall J. R.; Fennell C. J. Two-Dimensional Molecular Simulations Using Rose Potentials . J. Mol. Liq. 2017, 228 , 11–18. 10.1016/j.molliq.2016.10.016. [ CrossRef ] [ Google Scholar ]
  • Maréchal Y. In The Hydrogen Bond and the Water Molecule. The Physics and Chemistry of Water, Aqueous and Bio Media. ; Maréchal Y., Ed.; Elsevier: Amsterdam, 2007. [ Google Scholar ]
  • Kumar P.; Buldyrev S. V.; Stanley H. E. A Tetrahedral Entropy for Water . Proc. Natl. Acad. Sci. U. S. A. 2009, 106 , 22130–22134. 10.1073/pnas.0911094106. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Shadrack Jabes B.; Nayar D.; Dhabal D.; Molinero V.; Chakravarty C. Water and Other Tetrahedral Liquids: Order, Anomalies and Solvation . J. Phys.: Condens. Matter 2012, 24 , 284116. 10.1088/0953-8984/24/28/284116. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Pamuk B.; Allen P. B.; Fernández-Serra M.-V. Electronic and Nuclear Quantum Effects on the Ice XI/ice Ih Phase Transition . Phys. Rev. B: Condens. Matter Mater. Phys. 2015, 92 , 134105. 10.1103/PhysRevB.92.134105. [ CrossRef ] [ Google Scholar ]
  • McKenzie R. H.; Bekker C.; Athokpam B.; Ramesh S. G. Effect of quantum nuclear motion on hydrogen bonding . J. Chem. Phys. 2014, 140 , 174508. 10.1063/1.4873352. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Habershon S.; Markland T. E.; Manolopoulos D. E. Competing quantum effects in the dynamics of a flexible water model . J. Chem. Phys. 2009, 131 , 024501. 10.1063/1.3167790. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Li X.-Z.; Walker B.; Michaelides A. Quantum nature of the hydrogen bond . Proc. Natl. Acad. Sci. U. S. A. 2011, 108 , 6369–6373. 10.1073/pnas.1016653108. [ CrossRef ] [ Google Scholar ]
  • Zeidler A.; Salmon P. S.; Fischer H. E.; Neuefeind J. C.; Simonson J. M.; Lemmel H.; Rauch H.; Markland T. E. Oxygen as a Site Specific Probe of the Structure of Water and Oxide Materials . Phys. Rev. Lett. 2011, 107 , 145501. 10.1103/PhysRevLett.107.145501. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Markland T. E.; Berne B. J. Unraveling quantum mechanical effects in water using isotopic fractionation . Proc. Natl. Acad. Sci. U. S. A. 2012, 109 , 7988–7991. 10.1073/pnas.1203365109. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Romanelli G.; Ceriotti M.; Manolopoulos D. E.; Pantalei C.; Senesi R.; Andreani C. Direct Measurement of Competing Quantum Effects on the Kinetic Energy of Heavy Water upon Melting . J. Phys. Chem. Lett. 2013, 4 , 3251–3256. 10.1021/jz401538r. [ CrossRef ] [ Google Scholar ]
  • Lamb D.; Scott W. D. Linear Growth Rates of Ice Crystals Grown From the Vapor Phase . J. Cryst. Growth 1972, 12 , 21–31. 10.1016/0022-0248(72)90333-8. [ CrossRef ] [ Google Scholar ]
  • Shultz M. J.; Bisson P. J.; Brumberg A. Best Face Forward: Crystal-Face Competition at the Ice-Water Interface . J. Phys. Chem. B 2014, 118 , 7972–7980. 10.1021/jp500956w. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Bentley W. A. Annual Summary of the Monthly Weather Review for 1902 ; Government Printing Office: Washington, DC, 1903. [ Google Scholar ]
  • Bentley W. A.; Humphreys W. J.. Snow Crystals , 1st ed; McGraw-Hill: New York, 1931. [ Google Scholar ]
  • Furukawa Y.; Wettlaufer J. S. Snow and Ice Crystals . Phys. Today 2007, 60 ( 12 ), 70–71. 10.1063/1.2825081. [ CrossRef ] [ Google Scholar ]
  • Pauling L. The Structure and Entropy of Ice and of Other Crystals With Some Randomness of Atomic Arrangement . J. Am. Chem. Soc. 1935, 57 , 2680–2684. 10.1021/ja01315a102. [ CrossRef ] [ Google Scholar ]
  • Vonnegut K. Cat’s Cradle ; Holt, Rinehart and Winston: New York, 1963. [ Google Scholar ]
  • Johari G.; Hallbrucker A.; Mayer E. The glass-liquid transition of hyperquenched water . Nature 1987, 330 , 552–553. 10.1038/330552a0. [ CrossRef ] [ Google Scholar ]
  • Hallbrucker A.; Mayer E. Calorimetric study of the vitrified liquid water to cubic ice phase transition . J. Phys. Chem. 1987, 91 , 503–505. 10.1021/j100287a002. [ CrossRef ] [ Google Scholar ]
  • Hallbrucker A.; Mayer E.; Johari G. The heat capacity and glass transition of hyperquenched glassy water . Philos. Mag. B 1989, 60 , 179–187. 10.1080/13642818908211189. [ CrossRef ] [ Google Scholar ]
  • Sellberg J. A.; Huang C.; McQueen T. A.; Loh N. D.; Laksmono H.; Schlesinger D.; Sierra R. G.; Nordlund D.; Hampton C. Y.; Starodub D.; et al. Ultrafast X-ray probing of water structure below the homogeneous ice nucleation temperature . Nature 2014, 510 , 381–384. 10.1038/nature13266. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Speedy R. J.; Angell C. A. Isothermal Compressibility of Supercooled Water and Evidence for a Thermodynamic Singularity at −45 °C . J. Chem. Phys. 1976, 65 , 851–858. 10.1063/1.433153. [ CrossRef ] [ Google Scholar ]
  • Limmer D. T.; Chandler D. The Putative Liquid-Liquid Transition Is a Liquid-Solid Transition in Atomistic Models of Water . J. Chem. Phys. 2011, 135 , 134503. 10.1063/1.3643333. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Smith J. D.; Cappa C. D.; Wilson K. R.; Cohen R. C.; Geissler P. L.; Saykally R. J. Unified Description of Temperature-Dependent Hydrogen-Bond Rearrangements in Liquid Water . Proc. Natl. Acad. Sci. U. S. A. 2005, 102 , 14171–14174. 10.1073/pnas.0506899102. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Geissler P. L. Temperature Dependence of Inhomogeneous Broadening: on the Meaning of Isosbestic Points . J. Am. Chem. Soc. 2005, 127 , 14930–14935. 10.1021/ja0545214. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Bosio L.; Chen S.; Teixeira J. Isochoric Temperature Differential of the x-ray structure factor and structural rearrangements in low-temperature heavy water . Phys. Rev. A: At., Mol., Opt. Phys. 1983, 27 , 1468. 10.1103/PhysRevA.27.1468. [ CrossRef ] [ Google Scholar ]
  • Nilsson A.; Pettersson L. G. M. The Structural Origin of Anomalous Properties of Liquid Water . Nat. Commun. 2015, 6 , 8998. 10.1038/ncomms9998. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Huang C.; et al. The Inhomogeneous Structure of Water at Ambient Conditions . Proc. Natl. Acad. Sci. U. S. A. 2009, 106 , 15214–15218. 10.1073/pnas.0904743106. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Debenedetti P. G. Metastable Liquids: Concepts and Principles ; Princeton University Press: 1996. [ Google Scholar ]
  • Sciortino F.; Geiger A.; Stanley E. Isochoric Differential Scattering Functions in Liquid water: the fifth neighbor as a network defect . Phys. Rev. Lett. 1990, 65 , 3452. 10.1103/PhysRevLett.65.3452. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Poole P. H.; Sciortino F.; Essmann U.; Stanley H. E. Phase Behaviour of Metastable Water . Nature 1992, 360 , 324–328. 10.1038/360324a0. [ CrossRef ] [ Google Scholar ]
  • Urquidi J.; Singh S.; Cho C. H.; Robinson G. W. Origin of Temperature and Pressure Effects on the Radial distribution function of water . Phys. Rev. Lett. 1999, 83 , 2348. 10.1103/PhysRevLett.83.2348. [ CrossRef ] [ Google Scholar ]
  • Debenedetti P. G. Supercooled and Glassy Water . J. Phys.: Condens. Matter 2003, 15 , R1669. 10.1088/0953-8984/15/45/R01. [ CrossRef ] [ Google Scholar ]
  • Azouzi M. E. M.; Ramboz C.; Lenain J.-F.; Caupin F. A coherent picture of water at extreme negative pressure . Nat. Phys. 2013, 9 , 38. 10.1038/nphys2475. [ CrossRef ] [ Google Scholar ]
  • Amann-Winkel K.; Gainaru C.; Handle P. H.; Seidl M.; Nelson H.; Böhmer R.; Loerting T. Waters second glass transition . Proc. Natl. Acad. Sci. U. S. A. 2013, 110 , 17720–17725. 10.1073/pnas.1311718110. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Pallares G.; El Mekki Azouzi M.; González M. A.; Aragones J. L.; Abascal J. L.; Valeriani C.; Caupin F. Anomalies in bulk supercooled water at negative pressure . Proc. Natl. Acad. Sci. U. S. A. 2014, 111 , 7936–7941. 10.1073/pnas.1323366111. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Abascal J. L. F.; Vega C. Widom Line and the Liquid-liquid Critical Point for the TIP4P/2005 Water Model . J. Chem. Phys. 2010, 133 , 234502. 10.1063/1.3506860. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Wikfeldt K. T.; Nilsson A.; Pettersson L. G. M. Spatially Inhomogeneous Bimodal Inherent Structure of Simulated Liquid Water . Phys. Chem. Chem. Phys. 2011, 13 , 19918–19924. 10.1039/c1cp22076d. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Santra B.; DiStasio R. A. Jr.; Martelli F.; Car R. Local Structure Analysis in Ab Initio Liquid Water . Mol. Phys. 2015, 113 , 2829–2841. 10.1080/00268976.2015.1058432. [ CrossRef ] [ Google Scholar ]
  • Errington J. R.; Debenedetti P. G. Relationship Between Structural Order and the Anomalies of Liquid Water . Nature 2001, 409 , 318–321. 10.1038/35053024. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • English N. J.; Tse J. S. Density Fluctuations in Liquid Water . Phys. Rev. Lett. 2011, 106 , 037801. 10.1103/PhysRevLett.106.037801. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Mallamace F.; Corsaro C.; Stanley H. E. Possible Relation of Water Structural Relaxation to Water Anomalies . Proc. Natl. Acad. Sci. U. S. A. 2013, 110 , 4899–4904. 10.1073/pnas.1221805110. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Sahle C. J.; Sternemann C.; Schmidt C.; Lehtola S.; Jahn S.; Simonelli L.; Huotari S.; Hakala M.; Pylkkänen T.; Nyrow A.; et al. Microscopic Structure of Water at Elevated Pressures and Temperatures . Proc. Natl. Acad. Sci. U. S. A. 2013, 110 , 6301–6306. 10.1073/pnas.1220301110. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Holten V.; Palmer J. C.; Poole P. H.; Debenedetti P. G.; Anisimov M. A. Two-State Thermodynamics of the ST2 Model for Supercooled Water . J. Chem. Phys. 2014, 140 , 104502. 10.1063/1.4867287. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Palmer J. C.; Martelli F.; Liu Y.; Car R.; Panagiotopoulos A. Z.; Debenedetti P. G. Metastable liquid-liquid transition in a molecular model of water . Nature 2014, 510 , 385. 10.1038/nature13405. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Ni Y.; Skinner J. Evidence for a liquid-liquid critical point in supercooled water within the E3B3 model and a possible interpretation of the kink in the homogeneous nucleation line . J. Chem. Phys. 2016, 144 , 214501. 10.1063/1.4952991. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Singh R. S.; Biddle J. W.; Debenedetti P. G.; Anisimov M. A. Two-state thermodynamics and the possibility of a liquid-liquid phase transition in supercooled TIP4P/2005 water . J. Chem. Phys. 2016, 144 , 144504. 10.1063/1.4944986. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Smallenburg F.; Sciortino F. Tuning the liquid-liquid transition by modulating the hydrogen-bond angular flexibility in a model for water . Phys. Rev. Lett. 2015, 115 , 015701. 10.1103/PhysRevLett.115.015701. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Reynolds J. A.; Gilbert D. B.; Tanford C. Empirical Correlation Between Hydrophobic Free Energy and Aqueous Cavity Surface Area . Proc. Natl. Acad. Sci. U. S. A. 1974, 71 , 2925–2927. 10.1073/pnas.71.8.2925. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Tanford C. The Hydrophobic Effect: Formation of Micelles and Biological Membranes , 2nd ed.; John Wiley: New York, 1980. [ Google Scholar ]
  • Dill K. A. Additivity Principles in Biochemistry . J. Biol. Chem. 1997, 272 , 701–704. 10.1074/jbc.272.2.701. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Fennell C. J.; Dill K. A. Physical Modeling of Aqueous Solvation . J. Stat. Phys. 2011, 145 , 209–226. 10.1007/s10955-011-0232-9. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Ben-Naim A. Thermodynamics of Aqueous Solutions of Noble Gases. I . J. Phys. Chem. 1965, 69 , 3240–3245. 10.1021/j100894a003. [ CrossRef ] [ Google Scholar ]
  • Ben-Naim A. A Simple Model for Demonstrating the Relation Between Solubility, Hydrophobic Interaction, and Structural Changes in the Solvent . J. Phys. Chem. 1978, 82 , 874–885. 10.1021/j100497a007. [ CrossRef ] [ Google Scholar ]
  • Ben-Naim A. Molecular Theory of Solutions ; Oxford University Press: 2006. [ Google Scholar ]
  • Ben-Amotz D. Water-Mediated Hydrophobic Interactions . Annu. Rev. Phys. Chem. 2016, 67 , 617–638. 10.1146/annurev-physchem-040215-112412. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Roseman M.; Jencks W. P. Interactions of Urea and Other Polar Compounds in Water . J. Am. Chem. Soc. 1975, 97 , 631–640. 10.1021/ja00836a027. [ CrossRef ] [ Google Scholar ]
  • Özal T. A.; van der Vegt N. F. Confusing Cause and Effect: Energy- Entropy Compensation in the Preferential Solvation of a Nonpolar Solute in Dimethyl Sulfoxide/Water Mixtures . J. Phys. Chem. B 2006, 110 , 12104–12112. 10.1021/jp061608i. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Lee M.-E.; Van der Vegt N. F. Molecular Thermodynamics of Methane Solvation in Tert-Butanol- Water Mixtures . J. Chem. Theory Comput. 2007, 3 , 194–200. 10.1021/ct600226h. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Snyder P. W.; Lockett M. R.; Moustakas D. T.; Whitesides G. M. Is It the Shape of the Cavity, or the Shape of the Water in the Cavity? . Eur. Phys. J.: Spec. Top. 2014, 223 , 853–891. 10.1140/epjst/e2013-01818-y. [ CrossRef ] [ Google Scholar ]
  • Yu H.-A.; Karplus M. A Thermodynamic Analysis of Solvation . J. Chem. Phys. 1988, 89 , 2366–2379. 10.1063/1.455080. [ CrossRef ] [ Google Scholar ]
  • Ben-Amotz D.; Underwood R. Unraveling Water’s Entropic Mysteries: A Unified View of Nonpolar, Polar, and Ionic Hydration . Acc. Chem. Res. 2008, 41 , 957–967. 10.1021/ar7001478. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Breiten B.; Lockett M. R.; Sherman W.; Fujita S.; Al-Sayah M.; Lange H.; Bowers C. M.; Heroux A.; Krilov G.; Whitesides G. M. Water Networks Contribute to Enthalpy/Entropy Compensation in Protein-Ligand Binding . J. Am. Chem. Soc. 2013, 135 , 15579–15584. 10.1021/ja4075776. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Makhatadze G. I.; Privalov P. L. Energetics of Interactions of Aromatic Hydrocarbons With . Biophys. Chem. 1994, 50 , 285–291. 10.1016/0301-4622(93)E0096-N. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Schravendijk P.; van der Vegt N. F. From Hydrophobic to Hydrophilic Solvation: An Application to Hydration of Benzene . J. Chem. Theory Comput. 2005, 1 , 643–652. 10.1021/ct049841c. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Frank H. S.; Evans M. W. Free Volume and Entropy in Condensed Systems III. Entropy in Binary Liquid Mixtures; Partial Molal Entropy in Dilute Solutions; Structure and Thermodynamics in Aqueous Electrolytes . J. Chem. Phys. 1945, 13 , 507–532. 10.1063/1.1723985. [ CrossRef ] [ Google Scholar ]
  • Pratt L. R.; Chandler D. Theory of the Hydrophobic Effect . J. Chem. Phys. 1977, 67 , 3683–3704. 10.1063/1.435308. [ CrossRef ] [ Google Scholar ]
  • Kronberg B. The Hydrophobic Effect . Curr. Opin. Colloid Interface Sci. 2016, 22 , 14–22. 10.1016/j.cocis.2016.02.001. [ CrossRef ] [ Google Scholar ]
  • Grdadolnik J.; Merzel F.; Avbelj F. Origin of Hydrophobicity and Enhanced Water Hydrogen Bond Strength Near Purely Hydrophobic Solutes . Proc. Natl. Acad. Sci. U. S. A. 2017, 114 , 322–327. 10.1073/pnas.1612480114. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Godec A.; Smith J. C.; Merzel F. Increase of Both Order and Disorder in the First Hydration Shell With Increasing Solute Polarity . Phys. Rev. Lett. 2011, 107 , 267801. 10.1103/PhysRevLett.107.267801. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Godec A.; Merzel F. Physical Origin Underlying the Entropy Loss Upon Hydrophobic Hydration . J. Am. Chem. Soc. 2012, 134 , 17574–17581. 10.1021/ja306464u. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Widom B.; Bhimalapuram P.; Koga K. The Hydrophobic Effect . Phys. Chem. Chem. Phys. 2003, 5 , 3085–3093. 10.1039/b304038k. [ CrossRef ] [ Google Scholar ]
  • Xu H.; Dill K. A. Water’s Hydrogen Bonds in the Hydrophobic Effect: A Simple Model . J. Phys. Chem. B 2005, 109 , 23611–23617. 10.1021/jp0526750. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Southall N. T.; Dill K. A.; Haymet A. A View of the Hydrophobic Effect . J. Phys. Chem. B 2002, 106 , 521–533. 10.1021/jp015514e. [ CrossRef ] [ Google Scholar ]
  • Chandler D. Interfaces and the Driving Force of Hydrophobic Assembly . Nature 2005, 437 , 640–647. 10.1038/nature04162. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Ben-Amotz D.; Widom B. Generalized Solvation Heat Capacities . J. Phys. Chem. B 2006, 110 , 19839–19849. 10.1021/jp061824r. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Hummer G.; Garde S.; Garcia A. E.; Pohorille A.; Pratt L. R. An Information Theory Model of Hydrophobic Interactions . Proc. Natl. Acad. Sci. U. S. A. 1996, 93 , 8951–8955. 10.1073/pnas.93.17.8951. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Patel A. J.; Varilly P.; Jamadagni S. N.; Hagan M. F.; Chandler D.; Garde S. Sitting at the Edge: How Biomolecules Use Hydrophobicity to Tune Their Interactions and Function . J. Phys. Chem. B 2012, 116 , 2498–2503. 10.1021/jp2107523. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Southall N. T.; Dill K. A. Potential of Mean Force Between Two Hydrophobic Solutes in Water . Biophys. Chem. 2002, 101–102 , 295–307. 10.1016/S0301-4622(02)00167-9. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Payne V. A.; Matubayasi N.; Murphy L. R.; Levy R. M. Monte Carlo Study of the Effect of Pressure on Hydrophobic Association . J. Phys. Chem. B 1997, 101 , 2054–2060. 10.1021/jp962977p. [ CrossRef ] [ Google Scholar ]
  • Hummer G.; Garde S.; Garcia A. E.; Paulaitis M. E.; Pratt L. R. The Pressure Dependence of Hydrophobic Interactions Is Consistent With the Observed Pressure Denaturation of Proteins . Proc. Natl. Acad. Sci. U. S. A. 1998, 95 , 1552–1555. 10.1073/pnas.95.4.1552. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Pratt L. R.; Pohorille A. Hydrophobic Effects and Modeling of Biophysical Aqueous Solution Interfaces . Chem. Rev. 2002, 102 , 2671–2692. 10.1021/cr000692+. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Hummer G.; Garde S.; Garcia A.; Paulaitis M. E.; Pratt L. R. Hydrophobic Effects on a Molecular Scale . J. Phys. Chem. B 1998, 102 , 10469–10482. 10.1021/jp982873+. [ CrossRef ] [ Google Scholar ]
  • Hummer G.; Garde S.; Garcıa A.; Pratt L. New Perspectives on Hydrophobic Effects . Chem. Phys. 2000, 258 , 349–370. 10.1016/S0301-0104(00)00115-4. [ CrossRef ] [ Google Scholar ]
  • Pratt L. R.; Chaudhari M. I.; Rempe S. B. Statistical Analyses of Hydrophobic Interactions: A Mini-Review . J. Phys. Chem. B 2016, 120 , 6455–6460. 10.1021/acs.jpcb.6b04082. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Sosso G. C.; Caravati S.; Rotskoff G.; Vaikuntanathan S.; Hassanali A. On the Role of Nonspherical Cavities in Short Length-Scale Density Fluctuations in Water . J. Phys. Chem. A 2017, 121 , 370–380. 10.1021/acs.jpca.6b11168. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Lum K.; Chandler D.; Weeks J. D. Hydrophobicity at Small and Large Length Scales . J. Phys. Chem. B 1999, 103 , 4570–4577. 10.1021/jp984327m. [ CrossRef ] [ Google Scholar ]
  • Stillinger F. H. The Physical Chemistry of Aqueous Systems ; Springer: 1973; pp 43–60. [ Google Scholar ]
  • Zhou R.; Berne B. J.; Germain R. The Free Energy Landscape for β Hairpin Folding in Explicit Water . Proc. Natl. Acad. Sci. U. S. A. 2001, 98 , 14931–14936. 10.1073/pnas.201543998. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Zhou R.; Berne B. J. Can a Continuum Solvent Model Reproduce the Free Energy Landscape of a β-Hairpin Folding in Water? . Proc. Natl. Acad. Sci. U. S. A. 2002, 99 , 12777–12782. 10.1073/pnas.142430099. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Zhou R.; Huang X.; Margulis C. J.; Berne B. J. Hydrophobic Collapse in Multidomain Protein Folding . Science 2004, 305 , 1605–1609. 10.1126/science.1101176. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Berne B. J.; Weeks J. D.; Zhou R. Dewetting and Hydrophobic Interaction in Physical and Biological Systems . Annu. Rev. Phys. Chem. 2009, 60 , 85–103. 10.1146/annurev.physchem.58.032806.104445. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Xi E.; Patel A. J. The Hydrophobic Effect, and Fluctuations: The Long and the Short of It . Proc. Natl. Acad. Sci. U. S. A. 2016, 113 , 4549–4551. 10.1073/pnas.1603014113. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Vaikuntanathan S.; Rotskoff G.; Hudson A.; Geissler P. L. Necessity of capillary modes in a minimal model of nanoscale hydrophobic solvation . Proc. Natl. Acad. Sci. U. S. A. 2016, 113 , E2224–E2230. 10.1073/pnas.1513659113. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Pitera J. W.; van Gunsteren W. F. The Importance of Solute-Solvent van der Waals Interactions With Interior Atoms of Biopolymers . J. Am. Chem. Soc. 2001, 123 , 3163–3164. 10.1021/ja0057474. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Wallqvist A.; Berne B. Computer Simulation of Hydrophobic Hydration Forces on Stacked Plates at Short Range . J. Phys. Chem. 1995, 99 , 2893–2899. 10.1021/j100009a053. [ CrossRef ] [ Google Scholar ]
  • Huang X.; Zhou R.; Berne B. J. Drying and Hydrophobic Collapse of Paraffin Plates . J. Phys. Chem. B 2005, 109 , 3546–3552. 10.1021/jp045520l. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Hua L.; Huang X.; Zhou R.; Berne B. J. Dynamics of Water Confined in the Interdomain Region of a Multidomain Protein . J. Phys. Chem. B 2006, 110 , 3704–3711. 10.1021/jp055399y. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Biedermann F.; Nau W. M.; Schneider H.-J. The Hydrophobic Effect Revisitedtudies With Supramolecular Complexes Imply High-Energy Water as a Noncovalent Driving Force . Angew. Chem., Int. Ed. 2014, 53 , 11158–11171. 10.1002/anie.201310958. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Hillyer M. B.; Gibb B. C. Molecular Shape and the Hydrophobic Effect . Annu. Rev. Phys. Chem. 2016, 67 , 307–329. 10.1146/annurev-physchem-040215-112316. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Setny P.; Baron R.; Kekenes-Huskey P. M.; McCammon J. A.; Dzubiella J. Solvent Fluctuations in Hydrophobic Cavity-Ligand Binding Kinetics . Proc. Natl. Acad. Sci. U. S. A. 2013, 110 , 1197–1202. 10.1073/pnas.1221231110. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Bellissent-Funel M.-C.; Hassanali A.; Havenith M.; Henchman R.; Pohl P.; Sterpone F.; van der Spoel D.; Xu Y.; Garcia A. E. Water Determines the Structure and Dynamics of Proteins . Chem. Rev. 2016, 116 , 7673–7697. 10.1021/acs.chemrev.5b00664. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Patel A. J.; Garde S. Efficient Method to Characterize the Context-Dependent Hydrophobicity of Proteins . J. Phys. Chem. B 2014, 118 , 1564–1573. 10.1021/jp4081977. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Patel A. J.; Varilly P.; Jamadagni S. N.; Acharya H.; Garde S.; Chandler D. Extended Surfaces Modulate Hydrophobic Interactions of Neighboring Solutes . Proc. Natl. Acad. Sci. U. S. A. 2011, 108 , 17678–17683. 10.1073/pnas.1110703108. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Kanduc M.; Schlaich A.; Schneck E.; Netz R. R. Water-Mediated Interactions Between Hydrophilic and Hydrophobic Surfaces . Langmuir 2016, 32 , 8767–8782. 10.1021/acs.langmuir.6b01727. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Ben-Amotz D. Hydrophobic Ambivalence: Teetering on the Edge of Randomness . J. Phys. Chem. Lett. 2015, 6 , 1696–1701. 10.1021/acs.jpclett.5b00404. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Jabes B. S.; Bratko D.; Luzar A. Universal Repulsive Contribution to the Solvent-Induced Interaction Between Sizable, Curved Hydrophobes . J. Phys. Chem. Lett. 2016, 7 , 3158–3163. 10.1021/acs.jpclett.6b01442. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Marcus Y. Effect of Ions on the Structure of Water: Structure Making and Breaking . Chem. Rev. 2009, 109 , 1346–1370. 10.1021/cr8003828. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Gurney R. W. Ionic Processes in Solution ; McGraw-Hill: New York, 1953. [ Google Scholar ]
  • Krestov G. A. Thermodynamics of Solvation ; Ellis Harwood: New York, 1990. [ Google Scholar ]
  • Samoilov O. Y. A New Approach to the Study of Hydration of Ions in Aqueous Solutions . Discuss. Faraday Soc. 1957, 24 , 141–146. 10.1039/df9572400141. [ CrossRef ] [ Google Scholar ]
  • Ohtaki H.; Radnai T. Structure and Dynamics of Hydrated Ions . Chem. Rev. 1993, 93 , 1157–1204. 10.1021/cr00019a014. [ CrossRef ] [ Google Scholar ]
  • Hunenberger P.; Reif M.. Single-Ion Solvation. Experimental and Theoretical Approaches to Elusive Thermodynamic Quantities ; RSC Publishing: Cambridge, U.K., 2011. [ Google Scholar ]
  • Kondoh M.; Ohshima Y.; Tsubouchi M. Ion Effects on the Structure of Water Studied by Terahertz Time-Domain Spectroscopy . Chem. Phys. Lett. 2014, 591 , 317–322. 10.1016/j.cplett.2013.11.055. [ CrossRef ] [ Google Scholar ]
  • Hummer G.; Pratt L. R.; Garcia A. E. Free Energy of Ionic Hydration . J. Phys. Chem. 1996, 100 , 1206–1215. 10.1021/jp951011v. [ CrossRef ] [ Google Scholar ]
  • Yang L.; Fan Y.; Gao Y. Q. Differences of Cations and Anions: Their Hydration, Surface Adsorption, and Impact on Water Dynamics . J. Phys. Chem. B 2011, 115 , 12456–12465. 10.1021/jp207652h. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Collins K. D. Charge Density-Dependent Strength of Hydration and Biological Structure . Biophys. J. 1997, 72 , 65–76. 10.1016/S0006-3495(97)78647-8. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • McDevit W. F.; Long F. A. The Activity Coefficient of Benzene in Aqueous Salt Solutions . J. Am. Chem. Soc. 1952, 74 , 1773–1777. 10.1021/ja01127a048. [ CrossRef ] [ Google Scholar ]
  • Von Hippel P. H.; Schleich T. Ion Effects on the Solution Structure of Biological Macromolecules . Acc. Chem. Res. 1969, 2 , 257–265. 10.1021/ar50021a001. [ CrossRef ] [ Google Scholar ]
  • Baldwin R. L. How Hofmeister Ion Interactions Affect Protein Stability . Biophys. J. 1996, 71 , 2056–2063. 10.1016/S0006-3495(96)79404-3. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Hofmeister F. Ueber die Constitution des flüssigen Wassers . Naunyn-Schmiedeberg’s Arch. Pharmacol. 1888, 24 , 247–260. 10.1007/BF01918191. [ CrossRef ] [ Google Scholar ]
  • Kunz W.; Henle J.; Ninham B. W. ‘Zur Lehre von der Wirkungen der Salze’ (About the Science of the Effect of Salts): Franz Hofmeister’s Historical Papers . Curr. Opin. Colloid Interface Sci. 2004, 9 , 19–37. 10.1016/j.cocis.2004.05.005. [ CrossRef ] [ Google Scholar ]
  • Wen-Hui X.; Jing-Zhe S.; Xi-Ming X. Studies on the Activity Coefficient of Benzene and Its Derivatives in Aqueous Salt Solutions . Thermochim. Acta 1990, 169 , 271–286. 10.1016/0040-6031(90)80154-Q. [ CrossRef ] [ Google Scholar ]
  • Long F. A.; McDevit W. F. Activity Coefficients of Nonelectrolyte Solutes in Aqueous Salt Solutions . Chem. Rev. 1952, 51 , 119–169. 10.1021/cr60158a004. [ CrossRef ] [ Google Scholar ]
  • Falabella J. B.; Teja A. S. Henry’s Constants of 2-Ketones in Aqueous Solutions Containing Inorganic or Quaternary Ammonium Salts . Ind. Eng. Chem. Res. 2008, 47 , 4505–4509. 10.1021/ie8000327. [ CrossRef ] [ Google Scholar ]
  • Wen W.-Y.; Hung J. H. Thermodynamics of Hydrocarbon Gases in Aqueous Tetraalkylammonium Salt Solutions . J. Phys. Chem. 1970, 74 , 170–180. 10.1021/j100696a032. [ CrossRef ] [ Google Scholar ]
  • Perez-Tejeda P.; Maestre A.; Delgado-Cobos P.; Burgess J. Single-Ion Setschenow Coefficients for Several Hydrophobic Non-Electrolytes in Aqueous Electrolyte Solutions . Can. J. Chem. 1990, 68 , 243–246. 10.1139/v90-032. [ CrossRef ] [ Google Scholar ]
  • Pegram L. M.; Record M. T. Jr. Thermodynamic Origin of Hofmeister Ion Effects . J. Phys. Chem. B 2008, 112 , 9428–9436. 10.1021/jp800816a. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Smith P. E. Computer Simulation of Cosolvent Effects on Hydrophobic Hydration . J. Phys. Chem. B 1999, 103 , 525–534. 10.1021/jp983303c. [ CrossRef ] [ Google Scholar ]
  • Kalra A.; Tugcu N.; Cramer S. M.; Garde S. Salting-In and Salting-Out of Hydrophobic Solutes in Aqueous Salt Solutions . J. Phys. Chem. B 2001, 105 , 6380–6386. 10.1021/jp010568+. [ CrossRef ] [ Google Scholar ]
  • Zhang Y.; Cremer P. S. Interactions Between Macromolecules and Ions: The Hofmeister Series . Curr. Opin. Chem. Biol. 2006, 10 , 658–663. 10.1016/j.cbpa.2006.09.020. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Fennell C. J.; Bizjak A.; Vlachy V.; Dill K. A. Ion Pairing in Molecular Simulations of Aqueous Alkali Halide Solutions . J. Phys. Chem. B 2009, 113 , 6782–6791. 10.1021/jp809782z. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Gujt J.; Bester-Rogac M.; Hribar-Lee B. An Investigation of Ion-Pairing of Alkali Metal Halides in Aqueous Solutions Using the Electrical Conductivity and the Monte Carlo Computer Simulation Methods . J. Mol. Liq. 2014, 190 , 34–41. 10.1016/j.molliq.2013.09.025. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Chorny I.; Dill K. A.; Jacobson M. P. Surfaces Affect Ion Pairing . J. Phys. Chem. B 2005, 109 , 24056–24060. 10.1021/jp055043m. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Head-Gordon T.; Sorenson J. M.; Pertsemlidis A.; Glaeser R. M. Differences in Hydration Structure Near Hydrophobic and Hydrophilic Amino Acids . Biophys. J. 1997, 73 , 2106–2115. 10.1016/S0006-3495(97)78241-9. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Dixit S.; Crain J.; Poon W. C. K.; Finney J. L.; Soper A. K. Molecular Segregation Observed in a Concentrated Alcohol-water Solution . Nature 2002, 416 , 829–832. 10.1038/416829a. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Chaplin M. F. In Adsorption and Phase Behaviour in Nanochannels and Nanotubes ; Dunne L., Manos G., Eds.; Springer: New York, 2010; pp 241–255. [ Google Scholar ]
  • Björneholm O.; Hansen M. H.; Hodgson A.; Liu L.-M.; Limmer D. T.; Michaelides A.; Pedevilla P.; Rossmeisl J.; Shen H.; Tocci G.; et al. Water at Interfaces . Chem. Rev. 2016, 116 , 7698–7726. 10.1021/acs.chemrev.6b00045. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Cicero G.; Grossman J. C.; Schwegler E.; Gygi F.; Galli G. Water Confined in Nanotubes and Between Graphene Sheets: A First Principle Study . J. Am. Chem. Soc. 2008, 130 , 1871–1878. 10.1021/ja074418+. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Agmon N.; Bakker H. J.; Campen R. K.; Henchman R. H.; Pohl P.; Roke S.; Thämer M.; Hassanali A. Protons and Hydroxide Ions in Aqueous Systems . Chem. Rev. 2016, 116 , 7642–7672. 10.1021/acs.chemrev.5b00736. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Stiopkin I. V.; Weeraman C.; Pieniazek P. A.; Shalhout F. Y.; Skinner J. L.; Benderskii A. V. Hydrogen Bonding at the Water Surface Revealed by Isotopic Dilution Spectroscopy . Nature 2011, 474 , 192–195. 10.1038/nature10173. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Petersen P. B.; Iyengar S. S.; Day T. J. F.; Voth G. A. The Hydrated Proton at the Water Liquid/Vapor Interface . J. Phys. Chem. B 2004, 108 , 14804–14806. 10.1021/jp046716o. [ CrossRef ] [ Google Scholar ]
  • Petersen P. B.; Saykally R. J. Evidence for an Enhanced Hydronium Concentration at the Liquid Water Surface . J. Phys. Chem. B 2005, 109 , 7976–7980. 10.1021/jp044479j. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Buch V.; Milet A.; Vacha R.; Jungwirth P.; Devlin J. P. Water Surface Is Acidic . Proc. Natl. Acad. Sci. U. S. A. 2007, 104 , 7342–7347. 10.1073/pnas.0611285104. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Hub J. S.; Wolf M. G.; Caleman C.; van Maaren P. J.; Groenhof G.; van der Spoel D. Thermodynamics of Hydronium and Hydroxide Surface Solvation . Chem. Sci. 2014, 5 , 1745–1749. 10.1039/c3sc52862f. [ CrossRef ] [ Google Scholar ]
  • Jungwirth P.; Tobias D. J. Specific Ion Effects at the Air/Water Interface . Chem. Rev. 2006, 106 , 1259–1281. 10.1021/cr0403741. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Otten D. E.; Shaffer P. R.; Geissler P. L.; Saykally R. J. Elucidating the Mechanism of Selective Ion Adsorption to the Liquid Water Surface . Proc. Natl. Acad. Sci. U. S. A. 2012, 109 , 701–705. 10.1073/pnas.1116169109. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Caleman C.; Hub J. S.; van Maaren P. J.; van der Spoel D. Atomistic Simulation of Ion Solvation in Water Explains Surface Preference of Halides . Proc. Natl. Acad. Sci. U. S. A. 2011, 108 , 6838–6842. 10.1073/pnas.1017903108. [ CrossRef ] [ Google Scholar ]
  • Huang Z.; Hua W.; Verreault D.; Allen H. C. Salty Glycerol Versus Salty Water Surface Organization: Bromide and Iodide Surface Propensities . J. Phys. Chem. A 2013, 117 , 6346–6353. 10.1021/jp4020228. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Hummer G.; Rasaiah J. C.; Noworyta J. P. Water Conduction Through the Hydrophobic Channel of a Carbon Nanotube . Nature 2001, 414 , 188–190. 10.1038/35102535. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Holt J. K.; Park H. G.; Wang Y.; Stadermann M.; Artyukhin A. B.; Grigoropoulos C. P.; Noy A.; Bakajin O. Fast Mass Transport Through Sub-2-Nanometer Carbon Nanotubes . Science 2006, 312 , 1034–1037. 10.1126/science.1126298. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Naguib N.; Ye H.; Gogotsi Y.; Yazicioglu A. G.; Megaridis C. M.; Yoshimura M. Observation of Water Confined in Nanometer Channels of Closed Carbon Nanotubes . Nano Lett. 2004, 4 , 2237–2243. 10.1021/nl0484907. [ CrossRef ] [ Google Scholar ]
  • Major R. C.; Houston J. E.; McGrath M. J.; Siepmann J. I.; Zhu X.-Y. Viscous Water Meniscus Under Nanoconfinment . Phys. Rev. Lett. 2006, 96 , 177803. 10.1103/PhysRevLett.96.177803. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Agrawal K. V.; Shimizu S.; Drahushuk L. W.; Kilcoyne D.; Strano M. S. Observation of Extreme Phase Transition Temperatures of Water Confined Inside Isolated Carbon Nanotubes . Nat. Nanotechnol. 2017, 12 , 267–273. 10.1038/nnano.2016.254. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Cerveny S.; Mallamace F.; Swenson J.; Vogel M.; Xu L. Confined Water as Model of Supercooled Water . Chem. Rev. 2016, 116 , 7608–7625. 10.1021/acs.chemrev.5b00609. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Fornasiero F.; Park H. G.; Holt J. K.; Stadermann M.; Grigoropoulos C. P.; Noy A.; Bakajin O. Ion Exclusion by Sub-2-Nm Carbon Nanotube Pores . Proc. Natl. Acad. Sci. U. S. A. 2008, 105 , 17250–17255. 10.1073/pnas.0710437105. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Li H.; Francisco J. S.; Zeng X. C. Unraveling the Mechanism of Selective Ion Transport in Hydrophobic Subnanometer Channels . Proc. Natl. Acad. Sci. U. S. A. 2015, 112 , 10851–10856. 10.1073/pnas.1513718112. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Koga K.; Zeng X. C.; Tanaka H. Freezing of Confined Water: A Bilayer Ice Phase in Hydrophobic Nanopores . Phys. Rev. Lett. 1997, 79 , 5262–5265. 10.1103/PhysRevLett.79.5262. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Sanz E.; Vega C.; Abascal J. L. F.; MacDowell L. G. Phase Diagram of Water From Computer Simulation . Phys. Rev. Lett. 2004, 92 , 255701. 10.1103/PhysRevLett.92.255701. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Abascal J. L. F.; Sanz E.; García Fernández R.; Vega C. A Potential Model for the Study of Ices and Amorphous Water: TIP4P/Ice . J. Chem. Phys. 2005, 122 , 234511. 10.1063/1.1931662. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Báez L. A.; Clancy P. Phase Equilibria in Extended Simple Point Charge Ice-Water Systems . J. Chem. Phys. 1995, 103 , 9744–9755. 10.1063/1.469938. [ CrossRef ] [ Google Scholar ]
  • Koga K.; Gao G.; Tanaka H.; Zeng X. C. Formation of Ordered Ice Nanotubes Inside Carbon Nanotubes . Nature 2001, 412 , 802–805. 10.1038/35090532. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Takaiwa D.; Hatano I.; Koga K.; Tanaka H. Phase Diagram of Water in Carbon Nanotubes . Proc. Natl. Acad. Sci. U. S. A. 2008, 105 , 39–43. 10.1073/pnas.0707917105. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Maniwa Y.; Kataura H.; Abe M.; Udaka A.; Suzuki S.; Achiba Y.; Kira H.; Matsuda K.; Kadowaki H.; Okabe Y. Ordered Water Inside Carbon Nanotubes: Formation of Pentagonal to Octagonal Ice-Nanotubes . Chem. Phys. Lett. 2005, 401 , 534–538. 10.1016/j.cplett.2004.11.112. [ CrossRef ] [ Google Scholar ]
  • Ghosh S.; Ramanathan K. V.; Sood A. K. Water at Nanoscale Confined in Single-Walled Carbon Nanotubes Studied by NMR . Europhys. Lett. 2004, 65 , 678–684. 10.1209/epl/i2003-10160-9. [ CrossRef ] [ Google Scholar ]
  • Maheshwary S.; Patel N.; Sathyamurthy N.; Kulkarni A. D.; Gadre S. R. Structure and Stability of Water Clusters (H 2 O) n , N = 8 – 20: An Ab Initio Investigation . J. Phys. Chem. A 2001, 105 , 10525–10537. 10.1021/jp013141b. [ CrossRef ] [ Google Scholar ]
  • Chou I-M.; Sharma A.; Burruss R. C.; Shu J.; Mao H.; Hemley R. J.; Goncharov A. F.; Stern L. A.; Kirby S. H. Transformations in methane hydrates . Proc. Natl. Acad. Sci. U. S. A. 2000, 97 , 13484–13487. 10.1073/pnas.250466497. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Boswell R.; Collett T. S. Current perspectives on gas hydrate resources . Energy Environ. Sci. 2011, 4 , 1206–1215. 10.1039/C0EE00203H. [ CrossRef ] [ Google Scholar ]
  • Buch V.; Devlin J. P.; Monreal I. A.; Jagoda-Cwiklik B.; Uras-Aytemiz N.; Cwiklik L. Clathrate hydrates with hydrogen-bonding guests . Phys. Chem. Chem. Phys. 2009, 11 , 10245–10265. 10.1039/b911600c. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Devlin J. P. Catalytic activity of methanol in all-vapor subsecond clathrate-hydrate formation . J. Chem. Phys. 2014, 140 , 164505. 10.1063/1.4871879. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • McLaurin G.; Shin K.; Alavi S.; Ripmeester J. A. Antifreezes act as catalysts for methane hydrate formation from ice . Angew. Chem., Int. Ed. 2014, 53 , 10429–10433. 10.1002/anie.201403638. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Amtawong J.; Guo J.; Hale J. S.; Sengupta S.; Fleischer E. B.; Martin R. W.; Janda K. C. Propane Clathrate Hydrate Formation Accelerated by Methanol . J. Phys. Chem. Lett. 2016, 7 , 2346–2349. 10.1021/acs.jpclett.6b00982. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Wooldridge P. J.; Richardson H. H.; Devlin J. P. Mobile Bjerrum Defects: A Criterion for Ice-like Crystal Growth . J. Chem. Phys. 1987, 87 , 4126–4131. 10.1063/1.452917. [ CrossRef ] [ Google Scholar ]
  • Jacobson L. C.; Hujo W.; Molinero V. Thermodynamic Stability and Growth of Guest-Free Clathrate Hydrates: A Low-Density Crystal Phase of Water . J. Phys. Chem. B 2009, 113 , 10298–10307. 10.1021/jp903439a. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Falenty A.; Hansen T. C.; Kuhs W. F. Formation and Properties of Ice XVI Obtained by Emptying a Type sII Clathrate Hydrate . Nature 2014, 516 , 231–233. 10.1038/nature14014. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Debenedetti P.; Stillinger F. Supercooled Liquids and the Glass Transition . Nature 2001, 410 , 259. 10.1038/35065704. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Yue Y.-Z.; Angell C. A. Clarifying the Glass-Transition Behaviour Of water by comparison with hyperquenched inorganic glasses . Nature 2004, 427 , 717–720. 10.1038/nature02295. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Starr F. W.; Sastry S.; La Nave E.; Scala A.; Stanley H. E.; Sciortino F. Thermodynamic and Structural Aspects of The potential energy surface of simulated water . Phys. Rev. E: Stat. Phys., Plasmas, Fluids, Relat. Interdiscip. Top. 2001, 63 , 041201. 10.1103/PhysRevE.63.041201. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Gallo P.; Sciortino F.; Tartaglia P.; Chen S.-H. Slow Dynamics of Water Molecules in Supercooled States . Phys. Rev. Lett. 1996, 76 , 2730. 10.1103/PhysRevLett.76.2730. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Sciortino F.; Gallo P.; Tartaglia P.; Chen S. Supercooled Water and the Kinetic Glass Transition . Phys. Rev. E: Stat. Phys., Plasmas, Fluids, Relat. Interdiscip. Top. 1996, 54 , 6331. 10.1103/PhysRevE.54.6331. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Starr F. W.; Nielsen J. K.; Stanley E. Fast and Slow Dynamics of Hydrogen Bonds in Liquid Water . Phys. Rev. Lett. 1999, 82 , 2294. 10.1103/PhysRevLett.82.2294. [ CrossRef ] [ Google Scholar ]
  • Starr F. W.; Nielsen J. K.; Stanley E. Hydrogen-Bond Dynamics for the Extended Simple Point-Charge model of water . Phys. Rev. E: Stat. Phys., Plasmas, Fluids, Relat. Interdiscip. Top. 2000, 62 , 579. 10.1103/PhysRevE.62.579. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Schroder T. B.; Sastry S.; Dyre J.; Glotzer S. Crossover to Potential Energy Landscape Dominated Dynamics in a Model Glass Forming Liquid . J. Chem. Phys. 2000, 112 , 9834. 10.1063/1.481621. [ CrossRef ] [ Google Scholar ]
  • Giovambattista N.; Starr F. W.; Sciortino F.; Buldyrev S. V.; Stanley H. E. Transitions Between Inherent Structures in Water . Phys. Rev. E: Stat. Phys., Plasmas, Fluids, Relat. Interdiscip. Top. 2002, 65 , 041502. 10.1103/PhysRevE.65.041502. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Angell C. A. Formation of Glasses From Liquids and Biopolymers . Science 1995, 267 , 1924–1935. 10.1126/science.267.5206.1924. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Angell C. Relaxation in Liquids, Polymers and Plastic CrystalsStrong/Fragile Patterns and Problems . J. Non-Cryst. Solids 1991, 131–133 , 13–31. 10.1016/0022-3093(91)90266-9. [ CrossRef ] [ Google Scholar ]
  • Ito K.; Moynihan C. T.; Angell C. A. Thermodynamic Determination of Fragility in Liquids and a Fragile-To-Strong Liquid Transition in Water . Nature 1999, 398 , 492–495. 10.1038/19042. [ CrossRef ] [ Google Scholar ]
  • DeFries T.; Jonas J. Pressure Dependence of NMR Proton Spin-Lattice Relaxation Times and Shear Viscosity in Liquid Water in the Temperature Range-15–10° C . J. Chem. Phys. 1977, 66 , 896–901. 10.1063/1.433995. [ CrossRef ] [ Google Scholar ]
  • Singh L. P.; Issenmann B.; Caupin F. Pressure Dependence of Viscosity in Supercooled Water and a Unified Approach for Thermodynamic and Dynamic Anomalies of Water . Proc. Natl. Acad. Sci. U. S. A. 2017, 114 , 4312–4317. 10.1073/pnas.1619501114. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Prielmeier F.; Lang E.; Speedy R.; Lüdemann H.-D. The Pressure Dependence of Self Diffusion in Supercooled Light and Heavy Water . Ber. Bunsenges. Phys. Chem. 1988, 92 , 1111–1117. 10.1002/bbpc.198800282. [ CrossRef ] [ Google Scholar ]
  • Kumar A.; Kuchhal P.; Dass N.; Pathak P. Anomalous Behaviour in Sound Velocity of Water . Phys. Chem. Liq. 2011, 49 , 453–458. 10.1080/00319101003646538. [ CrossRef ] [ Google Scholar ]
  • Sciortino F.; Poole P. H.; Stanley H. E.; Havlin S. Lifetime of the Bond Network and Gel-Like Anomalies in Supercooled Water . Phys. Rev. Lett. 1990, 64 , 1686. 10.1103/PhysRevLett.64.1686. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Mishima O.; Stanley H. E. The Relationship Between Liquid, Supercooled and Glassy Water . Nature 1998, 396 , 329–335. 10.1038/24540. [ CrossRef ] [ Google Scholar ]
  • Natzle W. C.; Moore C. B. Recombination of Hydrogen Ion (H + ) and Hydroxide in Pure Liquid Water . J. Phys. Chem. 1985, 89 , 2605–2612. 10.1021/j100258a035. [ CrossRef ] [ Google Scholar ]
  • Reed C. A. Myths About the Proton. The Nature of H + in Condensed Media . Acc. Chem. Res. 2013, 46 , 2567–2575. 10.1021/ar400064q. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Markovitch O.; Agmon N. Structure and Energetics of the Hydronium Hydration Shells . J. Phys. Chem. A 2007, 111 , 2253–2256. 10.1021/jp068960g. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Mizuse K.; Fujii A.; Mikami N. Long Range Influence of an Excess Proton on the Architecture of the Hydrogen Bond Network in Large-Sized Water Clusters . J. Chem. Phys. 2007, 126 , 231101. 10.1063/1.2750669. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Eigen M.; de Maeyer L. Self-Dissociation and Protonic Charge Transport in Water and Ice . Proc. R. Soc. London, Ser. A 1958, 247 , 505–533. 10.1098/rspa.1958.0208. [ CrossRef ] [ Google Scholar ]
  • Eigen M. Proton Transfer, Acid-Base Catalysis, and Enzymatic Hydrolysis. Part I: Elementary Processes . Angew. Chem., Int. Ed. Engl. 1964, 3 , 1–19. 10.1002/anie.196400011. [ CrossRef ] [ Google Scholar ]
  • Zundel G. Hydration Structure and Intermolecular Interaction in Polyelectrolytes . Angew. Chem., Int. Ed. Engl. 1969, 8 , 499–509. 10.1002/anie.196904991. [ CrossRef ] [ Google Scholar ]
  • Silverstein T. P. The Aqueous Proton Is Hydrated by More Than One Water Molecule: Is the Hydronium Ion a Useful Conceit? . J. Chem. Educ. 2014, 91 , 608–610. 10.1021/ed400559t. [ CrossRef ] [ Google Scholar ]
  • Stoyanov E. S.; Stoyanova I. V.; Reed C. A. The Structure of the Hydrogen Ion (H aq + ) in Water . J. Am. Chem. Soc. 2010, 132 , 1484–1485. 10.1021/ja9101826. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Stoyanov E. S.; Stoyanova I. V.; Reed C. A. The Unique Nature of H + in Water . Chem. Sci. 2011, 2 , 462–472. 10.1039/C0SC00415D. [ CrossRef ] [ Google Scholar ]
  • Tuckerman M. E.; Marx D.; Parrinello M. The Nature and Transport Mechanism of Hydrated Hydroxide Ions in Aqueous Solution . Nature 2002, 417 , 925–929. 10.1038/nature00797. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Tuckerman M. E.; Chandra A.; Marx D. Structure and Dynamics of OH – (Aq) . Acc. Chem. Res. 2006, 39 , 151–158. 10.1021/ar040207n. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Marx D.; Chandra A.; Tuckerman M. E. Aqueous Basic Solutions: Hydroxide Solvation, Structural Diffusion, and Comparison to the Hydrated Proton . Chem. Rev. 2010, 110 , 2174–2216. 10.1021/cr900233f. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Pliego J. R.; Riveros J. M. Ab Initio Study of the Hydroxide Ion-water Clusters: An Accurate Determination of the Thermodynamic Properties for the Processes nH 2 O + OH – → HO – (H 2 O) n (N = 1–4) . J. Chem. Phys. 2000, 112 , 4045–4052. 10.1063/1.480954. [ CrossRef ] [ Google Scholar ]
  • Chaudhuri C.; Wang Y.-S.; Jiang J. C.; Lee Y. T.; Chang H.-C.; Niedner-Schatteburg G. Infrared Spectra and Isomeric Structures of Hydroxide Ion-Water Clusters OH – (H 2 O) 1–5 : A Comparison With H 3 O + (H 2 O) 1–5 . Mol. Phys. 2001, 99 , 1161–1173. 10.1080/00268970110046312. [ CrossRef ] [ Google Scholar ]
  • Robertson W. H.; Diken E. G.; Price E. A.; Shin J.-W.; Johnson M. A. Spectroscopic Determination of the OH – Solvation Shell in the OH – (H 2 O)n Clusters . Science 2003, 299 , 1367–1372. 10.1126/science.1080695. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Cukierman S. Et Tu, Grotthuss! And Other Unfinished Stories . Biochim. Biophys. Acta, Bioenerg. 2006, 1757 , 876–885. 10.1016/j.bbabio.2005.12.001. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Marx D. Proton Transfer 200 Years After Von Grotthuss: Insights From Ab Initio Simulations . ChemPhysChem 2006, 7 , 1848–1870. 10.1002/cphc.200600128. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Marx D.; Tuckerman M. E.; Hutter J.; Parrinello M. The Nature of the Hydrated Excess Proton in Water . Nature 1999, 397 , 601–604. 10.1038/17579. [ CrossRef ] [ Google Scholar ]
  • Tuckerman M.; Laasonen K.; Sprik M.; Parrinello M. Ab Initio Molecular Dynamics Simulation of the Solvation and Transport of H 3 O + and OH – Ions in Water . J. Phys. Chem. 1995, 99 , 5749–5752. 10.1021/j100016a003. [ CrossRef ] [ Google Scholar ]
  • Geissler P. L.; Dellago C.; Chandler D.; Hutter J.; Parrinello M. Autoionization in Liquid Water . Science 2001, 291 , 2121–2124. 10.1126/science.1056991. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Hassanali A.; Prakash M. K.; Eshet H.; Parrinello M. On the Recombination of Hydronium and Hydroxide Ions in Water . Proc. Natl. Acad. Sci. U. S. A. 2011, 108 , 20410–20415. 10.1073/pnas.1112486108. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Hassanali A.; Giberti F.; Cuny J.; Kühne T. D.; Parrinello M. Proton Transfer Through the Water Gossamer . Proc. Natl. Acad. Sci. U. S. A. 2013, 110 , 13723–13728. 10.1073/pnas.1306642110. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Codorniu-Hernández E.; Kusalik P. G. Probing the Mechanisms of Proton Transfer in Liquid Water . Proc. Natl. Acad. Sci. U. S. A. 2013, 110 , 13697–13698. 10.1073/pnas.1312350110. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Bai C.; Herzfeld J. Special Pairs Are Decisive in the Autoionization and Recombination of Water . J. Phys. Chem. B 2017, 121 , 4213–4219. 10.1021/acs.jpcb.7b02110. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Peng Y.; Swanson J. M. J.; Kang S.-g.; Zhou R.; Voth G. A. Hydrated Excess Protons Can Create Their Own Water Wires . J. Phys. Chem. B 2015, 119 , 9212–9218. 10.1021/jp5095118. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Cramer C. J.; Truhlar D. G. Implicit Solvation Models: Equilibria, Structure, Spectra, and Dynamics . Chem. Rev. 1999, 99 , 2161–2200. 10.1021/cr960149m. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Tsui V.; Case D. A. Theory and Applications of the Generalized Born Solvation Model in Macromolecular Simulations . Biopolymers 2000, 56 , 275–291. 10.1002/1097-0282(2000)56:4<275::AID-BIP10024>3.0.CO;2-E. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Simonson T. Macromolecular Electrostatics: Continuum Models and Their Growing Pains . Curr. Opin. Struct. Biol. 2001, 11 , 243–252. 10.1016/S0959-440X(00)00197-4. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Hassan S. A.; Mehler E. L. A Critical Analysis of Continuum Electrostatics: The Screened Coulomb Potential-Implicit Solvent Model and the Study of the Alanine Dipeptide and Discrimination of Misfolded Structures of Proteins . Proteins: Struct., Funct., Genet. 2002, 47 , 45–61. 10.1002/prot.10059. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Lee M. S.; Salsbury F. R.; Brooks C. L. Novel Generalized Born Methods . J. Chem. Phys. 2002, 116 , 10606–10614. 10.1063/1.1480013. [ CrossRef ] [ Google Scholar ]
  • Orozco M.; Luque F. J. Theoretical Methods for the Description of the Solvent Effect in Biomolecular Systems . Chem. Rev. 2000, 100 , 4187–4226. 10.1021/cr990052a. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Goncalves P. F. B.; Stassen H. Calculation of the Free Energy of Solvation From Molecular Dynamics Simulations . Pure Appl. Chem. 2004, 76 , 231–240. 10.1351/pac200476010231. [ CrossRef ] [ Google Scholar ]
  • Tomasi J.; Persico M. Molecular Interactions in Solution: An Overview of Methods Based on Continuous Distributions of the Sovent . Chem. Rev. 1994, 94 , 2027–2094. 10.1021/cr00031a013. [ CrossRef ] [ Google Scholar ]
  • Bashford D.; Case D. A. Generalized Born Models of Macromolecular Solvation Effects . Annu. Rev. Phys. Chem. 2000, 51 , 129–152. 10.1146/annurev.physchem.51.1.129. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Zhang H.; Tan T.; van der Spoel D. Generalized Born and Explicit Solvent Models for Free Energy Calculations in Organic Solvents: Cyclodextrin Dimerization . J. Chem. Theory Comput. 2015, 11 , 5103–5113. 10.1021/acs.jctc.5b00620. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Warwicker J.; Watson H. C. Calculation of the Electric Potential in the Active Site Cleft Due to α-Helix Dipoles . J. Mol. Biol. 1982, 157 , 671–679. 10.1016/0022-2836(82)90505-8. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Sharp K. A.; Honig B. Calculating Total Electrostatic Energies With the Nonlinear Poisson-Boltzmann Equation . J. Phys. Chem. 1990, 94 , 7684–7692. 10.1021/j100382a068. [ CrossRef ] [ Google Scholar ]
  • Davis M. E.; McCammon J. A. Electrostatics in Biomolecular Structure and Dynamics . Chem. Rev. 1990, 90 , 509–521. 10.1021/cr00101a005. [ CrossRef ] [ Google Scholar ]
  • Honig B.; Nicholls A. Classical Electrostatics in Biology and Chemistry . Science 1995, 268 , 1144–1149. 10.1126/science.7761829. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Baker N. A.; Sept D.; Joseph S.; Holst M. J.; McCammon J. A. Electrostatics of Nanosystems: Application to Microtubules and the Ribosome . Proc. Natl. Acad. Sci. U. S. A. 2001, 98 , 10037–10041. 10.1073/pnas.181342398. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Luo R.; David L.; Gilson M. K. Accelerated Poisson-Boltzmann Calculations for Static and Dynamic Systems . J. Comput. Chem. 2002, 23 , 1244–1253. 10.1002/jcc.10120. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Feig M.; Onufriev A.; Lee M. S.; Im W.; Case D. A.; Brooks C. L. III. Performance Comparison of Generalized Born and Poisson Methods in the Calculation of Electrostatic Solvation Energies for Protein Structures . J. Comput. Chem. 2004, 25 , 265–284. 10.1002/jcc.10378. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Still W. C.; Tempczyk A.; Hawley R. C.; Hendrickson T. Semianalytical Treatment of Solvation for Molecular Mechanics and Dynamics . J. Am. Chem. Soc. 1990, 112 , 6127–6129. 10.1021/ja00172a038. [ CrossRef ] [ Google Scholar ]
  • Onufriev A. V.; Aguilar B. Accuracy of Continuum Electrostatic Calculations Based on Three Common Dielectric Boundary Definitions . J. Theor. Comput. Chem. 2014, 13 , 1440006. 10.1142/S0219633614400069. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Onufriev A.; Case D. A.; Bashford D. Effective Born Radii in the Generalized Born Approximation: The Importance of Being Perfect . J. Comput. Chem. 2002, 23 , 1297–1304. 10.1002/jcc.10126. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Onufriev A.; Bashford D.; Case D. Modification of the Generalized Born Model Suitable for Macromolecules . J. Phys. Chem. B 2000, 104 , 3712–3720. 10.1021/jp994072s. [ CrossRef ] [ Google Scholar ]
  • Gallicchio E.; Levy R. M. AGBNP: An Analytic Implicit Solvent Model Suitable for Molecular Dynamics Simulations and High-Resolution Modeling . J. Comput. Chem. 2004, 25 , 479–499. 10.1002/jcc.10400. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Mongan J.; Simmerling C.; McCammon J. A.; Case D. A.; Onufriev A. Generalized Born Model With a Simple, Robust Molecular Volume Correction . J. Chem. Theory Comput. 2007, 3 , 156–169. 10.1021/ct600085e. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Onufriev A.; Bashford D.; Case D. A. Exploring Protein Native States and Large-Scale Conformational Changes With a Modified Generalized Born Model . Proteins: Struct., Funct., Genet. 2004, 55 , 383–394. 10.1002/prot.20033. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Nguyen H.; Roe D. R.; Simmerling C. Improved Generalized Born Solvent Model Parameters for Protein Simulations . J. Chem. Theory Comput. 2013, 9 , 2020–2034. 10.1021/ct3010485. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Nguyen H.; Maier J.; Huang H.; Perrone V.; Simmerling C. Folding Simulations for Proteins With Diverse Topologies Are Accessible in Days With a Physics-Based Force Field and Implicit Solvent . J. Am. Chem. Soc. 2014, 136 , 13959–13962. 10.1021/ja5032776. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Warshel A.; Papazyan A. Electrostatic Effects in Macromolecules: Fundamental Concepts and Practical Modeling . Curr. Opin. Struct. Biol. 1998, 8 , 211–217. 10.1016/S0959-440X(98)80041-9. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Ratkova E. L.; Palmer D. S.; Fedorov M. V. Solvation Thermodynamics of Organic Molecules by the Molecular Integral Equation Theory: Approaching Chemical Accuracy . Chem. Rev. 2015, 115 , 6312–6356. 10.1021/cr5000283. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Mobley D. L.; Barber A. E.; Fennell C. J.; Dill K. A. Charge Asymmetries in Hydration of Polar Solutes . J. Phys. Chem. B 2008, 112 , 2405–2414. 10.1021/jp709958f. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Chandler D.; Andersen H. C. Optimal Cluster Expansions for Classical Fluids. 2. Theory of Molecular Liquids . J. Chem. Phys. 1972, 57 , 1930–1937. 10.1063/1.1678513. [ CrossRef ] [ Google Scholar ]
  • Hirata F.; Rossky P. J. An Extended RISM Equation for Molecular Polar Fluids . Chem. Phys. Lett. 1981, 83 , 329–334. 10.1016/0009-2614(81)85474-7. [ CrossRef ] [ Google Scholar ]
  • Pettitt B. M.; Rossky P. J. Alkali Halides in Water: Ion-Solvent Correlations and Ion-Ion Potentials of Mean Force at Infinite Dilution . J. Chem. Phys. 1986, 84 , 5836–5844. 10.1063/1.449894. [ CrossRef ] [ Google Scholar ]
  • Montgomery Pettitt B.; Karplus M. The Structure of Water Surrounding a Peptide: A Theoretical Approach . Chem. Phys. Lett. 1987, 136 , 383–386. 10.1016/0009-2614(87)80271-3. [ CrossRef ] [ Google Scholar ]
  • Lee P. H.; Maggiora G. M. Solvation Thermodynamics of Polar Molecules in Aqueous Solution by the XRISM Method . J. Phys. Chem. 1993, 97 , 10175–10185. 10.1021/j100141a045. [ CrossRef ] [ Google Scholar ]
  • Beglov D.; Roux B. An Integral Equation to Describe the Solvation of Polar Molecules in Liquid Water . J. Phys. Chem. B 1997, 101 , 7821–7826. 10.1021/jp971083h. [ CrossRef ] [ Google Scholar ]
  • Kovalenko A. Molecular Theory of Solvation: Methodology Summary and Illustrations . Condens. Matter Phys. 2015, 18 , 32601. 10.5488/CMP.18.32601. [ CrossRef ] [ Google Scholar ]
  • Skyner R. E.; McDonagh J. L.; Groom C. R.; van Mourik T.; Mitchell J. B. O. A Review of Methods for the Calculation of Solute Free Energies and the Modelling of System in Solution . Phys. Chem. Chem. Phys. 2015, 17 , 6174–6191. 10.1039/C5CP00288E. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Giambaşu G. M.; Luchko T.; Herschlag H.; York D. M.; Case D. A. Ion Counting From Explicit-Solvent Simulations and 3D-RISM . Biophys. J. 2014, 106 , 883–894. 10.1016/j.bpj.2014.01.021. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Giambaşu G. M.; Gebala M. K.; Panteva M. T.; Luchko T.; Case D. A.; York D. M. Competitive Interaction of Monovalent Cations With DNA From 3D-RISM . Nucleic Acids Res. 2015, 43 , 8405–8415. 10.1093/nar/gkv830. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Nguyen H. T.; Pabit S. A.; Pollack L.; Case D. A. Extracting Water and Ion Distributions From Solution X-Ray Scattering Experiments . J. Chem. Phys. 2016, 144 , 214105. 10.1063/1.4953037. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Genheden S.; Luchko T.; Gusarov S.; Kovalenko A.; Ryde U. An MM/3D-RISM Approach for Ligand Binding Affinities . J. Phys. Chem. B 2010, 114 , 8505–8516. 10.1021/jp101461s. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Kiyota Y.; Yoshida N.; Hirata F. A New Approach for Investigating the Molecular Recognition of Protein: Toward Structure-Based Drug Design Based on the 3D-RISM Theory . J. Chem. Theory Comput. 2011, 7 , 3803–3815. 10.1021/ct200358h. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Nikolić D.; Blinov N.; Wishart D.; Kovalenko A. 3D-RISM-Dock: A New Fragment-Based Drug Design Protocol . J. Chem. Theory Comput. 2012, 8 , 3356–3372. 10.1021/ct300257v. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Yamazaki T.; Fenniri H.; Kovalenko A. Structural Water Drives Self-Assembly of Organic Rosette Nanotubes and Holds Host Atoms in the Channel . ChemPhysChem 2010, 11 , 361–367. 10.1002/cphc.200900324. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Drabik P.; Gusarov S.; Kovalenko A. Microtubule Stability Studied by Three-Dimensional Molecular Theory of Solvation . Biophys. J. 2007, 92 , 394–403. 10.1529/biophysj.106.089987. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Yamazaki T.; Blinov N.; Wishart D.; Kovalenko A. Hydration Effects on the HET-s Prion and Amyloid-Beta Fibrillous Aggregates, Studied With Three-Dimensional Molecular Theory of Solvation . Biophys. J. 2008, 95 , 4540–4548. 10.1529/biophysj.107.123000. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Misin M.; Fedorov M. V.; Palmer D. S. Communication: Accurate Hydration Free Energies at a Wide Range of Temperatures From 3D-RISM . J. Chem. Phys. 2015, 142 , 091105. 10.1063/1.4914315. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Joung I. S.; Luchko T.; Case D. A. Simple Electrolyte Solutions: Comparison of DRISM and Molecular Dynamics Results for Alkali Halide Solutions . J. Chem. Phys. 2013, 138 , 044103. 10.1063/1.4775743. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Giambasu G. M.; Luchko T.; Herschlag D.; York D. M.; Case D. A. Ion Counting From Explicit-Solvent Simulation and 3D-RISM . Biophys. J. 2014, 106 , 883–894. 10.1016/j.bpj.2014.01.021. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Truchon J.-F.; Pettitt B. M.; Labute P. A Cavity Corrected 3D-RISM Functional for Accurate Solvation Free Energies . J. Chem. Theory Comput. 2014, 10 , 934–941. 10.1021/ct4009359. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Luchko T.; Blinov N.; Limon G. C.; Joyce K. P.; Kovalenko A. SAMPL5:3D-RISM Partition Coefficient Calculations With Partial Molar Volume Corrections and Solute Conformational Sampling . J. Comput.-Aided Mol. Des. 2016, 30 , 1115–1127. 10.1007/s10822-016-9947-7. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Howard J. J.; Perkyns J. S.; Choudhury N.; Pettitt B. M. Integral Equation Study of the Hydrophobic Interaction Between Graphene Plates . J. Chem. Theory Comput. 2008, 4 , 1928–1939. 10.1021/ct8002817. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Misin M.; Vainikka P. A.; Fedorov M. V.; Palmer D. S. Salting-Out Effects by Pressure-Corrected 3D-RISM . J. Chem. Phys. 2016, 145 , 194501. 10.1063/1.4966973. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Johnson J.; Case D. A.; Yamazaki T.; Gusarov S.; Kovalenko A.; Luchko T. Small Molecule Hydration Energy and Entropy From 3D-RISM . J. Phys.: Condens. Matter 2016, 28 , 344002. 10.1088/0953-8984/28/34/344002. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Fennell C. J.; Kehoe C.; Dill K. A. Oil/Water Transfer Is Partly Driven by Molecular Shape, Not Just Size . J. Am. Chem. Soc. 2010, 132 , 234–240. 10.1021/ja906399e. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Fennell C. J.; Kehoe C. W.; Dill K. A. Modeling Aqueous Solvation With Semi-Explicit Assembly . Proc. Natl. Acad. Sci. U. S. A. 2011, 108 , 3234–3239. 10.1073/pnas.1017130108. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Li L.; Fennell C. J.; Dill K. A. Field-Sea: A Model for Computing the Solvation Free Energies of Nonpolar, Polar, and Charged Solutes in Water . J. Phys. Chem. B 2014, 118 , 6431–6437. 10.1021/jp4115139. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Kehoe C. W.; Fennell C. J.; Dill K. A. Testing the Semi-Explicit Assembly Solvation Model in the SAMPL3 Community Blind Test . J. Comput.-Aided Mol. Des. 2012, 26 , 563–568. 10.1007/s10822-011-9536-8. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Li L.; Dill K. A.; Fennell C. J. Testing the Semi-Explicit Assembly Model of Aqueous Solvation in the SAMPL4 Challenge . J. Comput.-Aided Mol. Des. 2014, 28 , 259–264. 10.1007/s10822-014-9712-8. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Brini E.; Paranahewage S. S.; Fennell C. J.; Dill K. A. Adapting the Semi-Explicit Assembly Solvation Model for Estimating Water-Cyclohexane Partitioning With the SAMPL5 Molecules . J. Comput.-Aided Mol. Des. 2016, 30 , 1067–1077. 10.1007/s10822-016-9961-9. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Li L.; Fennell C. J.; Dill K. A. Small Molecule Solvation Changes Due to the Presence of Salt Are Governed by the Cost of Solvent Cavity Formation and Dispersion . J. Chem. Phys. 2014, 141 , 22D518. 10.1063/1.4900890. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Lukšič M.; Fennell C. J.; Dill K. A. Using Interpolation for Fast and Accurate Calculation of Ion-Ion Interactions . J. Phys. Chem. B 2014, 118 , 8017–8025. 10.1021/jp501141j. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Kovalenko A.; Hirata F. Potentials of Mean Force of Simple Ions in Ambient Aqueous Solution. II. Solvation Structure From the Three-Dimensional Reference Interaction Site Model Approach, and Comparison With Simulations . J. Chem. Phys. 2000, 112 , 10403–10417. 10.1063/1.481677. [ CrossRef ] [ Google Scholar ]
  • Guthrie J. P. A Blind Challenge for Computational Solvation Free Energies: Introduction and Overview . J. Phys. Chem. B 2009, 113 , 4501–4507. 10.1021/jp806724u. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Geballe M. T.; Skillman A. G.; Nicholls A.; Guthrie J. P.; Taylor P. J. The SAMPL2 Blind Prediction Challenge: Introduction and Overview . J. Comput.-Aided Mol. Des. 2010, 24 , 259–279. 10.1007/s10822-010-9350-8. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Skillman A. G. SAMPL3: Blinded Prediction of Host-Guest Binding Affinities, Hydration Free Energies, and Trypsin Inhibitors . J. Comput.-Aided Mol. Des. 2012, 26 , 473–474. 10.1007/s10822-012-9580-z. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Mobley D. L.; Wymer K. L.; Lim N. M.; Guthrie J. P. Blind Prediction of Solvation Free Energies From the SAMPL4 Challenge . J. Comput.-Aided Mol. Des. 2014, 28 , 135–150. 10.1007/s10822-014-9718-2. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Bannan C. C.; Burley K. H.; Chiu M.; Shirts M. R.; Gilson M. K.; Mobley D. L. Blind Prediction of Cyclohexane-Water Distribution Coefficients From the SAMPL5 Challenge . J. Comput.-Aided Mol. Des. 2016, 30 , 927–944. 10.1007/s10822-016-9954-8. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]

Syllabus Edition

First teaching 2020

Last exams 2024

Water & the Hydrogen Bond ( CIE A Level Biology )

Revision note.

Lára

Biology Lead

Water Molecules: Hydrogen Bonds

  • Water is of great biological importance. It is the medium in which all metabolic reactions take place in cells. Between 70% to 95% of the mass of a cell is water
  • As 71% of the Earth’s surface is covered in water it is a major habitat for organisms
  • Water is composed of atoms of hydrogen and oxygen. One atom of oxygen combines with two atoms of hydrogen by sharing electrons ( covalent bonding )
  • The oxygen atom attracts the electrons more strongly than the hydrogen atoms, resulting in a weak negatively charged region on the oxygen atom (δ - ) and a weak positively charged region on the hydrogen atoms(δ + ), this also results in the asymmetrical shape
  • This separation of charge due to the electrons in the covalent bonds being unevenly shared is called a dipole . When a molecule has one end that is negatively charged and one end that is positively charged it is also a polar molecule
  • Water is a polar molecule

A water molecule, downloadable AS & A Level Biology revision notes

The covalent bonds of water make it a polar molecule

  • As a result of the polarity of water hydrogen bonds form between the positive and negatively charged regions of adjacent water molecules
  • Hydrogen bonds are weak, when there are few, so they are constantly breaking and reforming. However when there are large numbers present they form a strong structure
  • An excellent solvent – many substances can dissolve in water
  • A relatively high specific heat capacity
  • A relatively high latent heat of vaporisation
  • Water is less dense when a solid
  • Water has high surface tension and cohesion
  • It acts as a reagent

Hydrogen bonds between water molecules, downloadable AS & A Level Biology revision notes

The polarity of water molecules allows hydrogen bonds to form between adjacent water molecules

It is important to know where the hydrogen bonds form between water molecules (oxygen of one water molecule to the hydrogen atom of another).

You've read 0 of your 10 free revision notes

Get unlimited access.

to absolutely everything:

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

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

the (exam) results speak for themselves:

Did this page help you?

Author: Lára

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

IMAGES

  1. Diary of a Water Molecule

    water molecule essay

  2. The Water Molecule

    water molecule essay

  3. UNIT 10A

    water molecule essay

  4. Water

    water molecule essay

  5. WATER CHEMISTRY

    water molecule essay

  6. Chemical and Physical Properties of Water

    water molecule essay

COMMENTS

  1. Lesson summary: Water and life (article)

    Water is polar. Water molecules are polar, with partial positive charges on the hydrogens, a partial negative charge on the oxygen, and a bent overall structure. This is because oxygen is more electronegative, meaning that it is better than hydrogen at attracting electrons. Water is an excellent solvent.

  2. Hydrogen bonds in water (article)

    A water molecule consists of two hydrogen atoms bonded to an oxygen atom, and its overall structure is bent. This is because the oxygen atom, in addition to forming bonds with the hydrogen atoms, also carries two pairs of unshared electrons. All of the electron pairs—shared and unshared—repel each other.

  3. Water

    The water on the surface of Earth is found mainly in its oceans (97.25 percent) and polar ice caps and glaciers (2.05 percent), with the balance in freshwater lakes, rivers, and groundwater.As Earth's population grows and the demand for fresh water increases, water purification and recycling become increasingly important. Interestingly, the purity requirements of water for industrial use ...

  4. Water

    Water is an inorganic compound with the chemical formula H 2 O.It is a transparent, tasteless, odorless, [c] and nearly colorless chemical substance.It is the main constituent of Earth's hydrosphere and the fluids of all known living organisms (in which it acts as a solvent [19]).It is vital for all known forms of life, despite not providing food energy or organic micronutrients.

  5. Water structure, properties and some applications

    This review is structured as follows. Section 2 discusses the extraordinary and sometimes anomalous properties of water from thermodynamic and surface science points of view, as well as the two-state theory of water and peculiar applications of water properties. Section 3 presents water interactions with hydrophobic and ionic compounds. Sections 4 and 5 discuss the phenomena of water bridge ...

  6. The Structure of Water

    The bent shape of the water molecule is critical because the polar O-H bonds do not cancel one another and the molecule as a whole is polar. Figure 2 below illustrates the net polarity of the water molecule. The oxygen is the negative end of the molecule, while the area between the hydrogen atoms is the positive end of the molecule.

  7. Water structure and science

    Liquid water has importance as a solvent, a solute, a reactant, a catalyst, and a biomolecule, structuring proteins, nucleic acid, and cells and controlling our consciousness. H 2 O is the second most common molecule in the Universe (behind hydrogen, H 2 ), the most abundant solid material and fundamental to star formation.

  8. (PDF) Water: Structure and Properties

    Selected physical properties of water are given in. Table 1. To put these in context, comparison is made to the organic. solvents methanol and dimethyl ether, where one and two. of the hydrogen ...

  9. The Water Molecule

    Abstract. This chapter presents descriptions of an isolated water molecule based on experiment and theory. The description based on experiment involves measurements made on water vapour at sufficiently low pressures or high temperatures to ensure that interactions between molecules are largely absent. The description based on theory provides ...

  10. Water

    Water undergoes various types of chemical reactions. One of the most important chemical properties of water is its ability to behave as both an acid (a proton donor) and a base (a proton acceptor), the characteristic property of amphoteric substances. This behaviour is most clearly seen in the autoionization of water: H 2 O(l) + H 2 O(l) ⇌ H 3 O + (aq) + OH − (aq), where the (l) represents ...

  11. Biological Roles of Water: Why is water necessary for life?

    Water allows everything inside cells to have the right shape at the molecular level. As shape is critical for biochemical processes, this is also one of water's most important roles. Figure 2: Water impacts cell shape. Water creates pressure inside the cell that helps it maintain shape. In the hydrated cell (left), the water pushes outward ...

  12. Understanding the Structure and Function of Water at the Molecular

    2.2.1 Water Molecule. It is well known that water consists of two elements, hydrogen (H) and oxygen (O). Two hydrogen atoms and one oxygen atom form a V-shaped structure (Fig. 2.3), which constitutes the water molecule (H 2 O). The experimental determination of the OH bond length of the gaseous free water molecule at room temperature and pressure is 0.9572 Å (1 Å = 10 -10 m), and the HOH ...

  13. Water Molecule

    Properties of Water. Molecular weight: The total of all the atomic masses of a molecule is about 18.0151 g/m in water. State: Water has three states, liquid state at room temperature, solid-state ...

  14. Solvent properties of water (article)

    A solvent is simply a substance that can dissolve other molecules and compounds, which are known as solutes. A homogeneous mixture of solvent and solute is called a solution, and much of life's chemistry takes place in aqueous solutions, or solutions with water as the solvent. Because of its polarity and ability to form hydrogen bonds, water ...

  15. 2.2 Water

    It is because water is essential to life; even minute traces of it on another planet can indicate that life could or did exist on that planet. Water is one of the more abundant molecules in living cells and the one most critical to life as we know it. Approximately 60-70 percent of your body is made up of water.

  16. How Water's Properties Are Encoded in Its Molecular Structure and

    At this level, a water molecule is understood by solving the Schrödinger equation for 10 electrons in the molecule, obtaining the covalent bonding of two hydrogen atoms and one oxygen atom. The hybridization of the molecular orbitals is sp 3, which means that a water molecule has tetrahedral structure, even though the three atoms are coplanar ...

  17. The Properties of Water

    Water has many essential roles in living organisms due to its properties: The polarity of water molecules; The presence and number of hydrogen bonds between water molecules; Solvent. As water is a polar molecule many ions (e.g. sodium chloride) and covalently bonded polar substances (e.g. glucose) will dissolve in it. This allows chemical reactions to occur within cells (as the dissolved ...

  18. Water & the Hydrogen Bond

    Water Molecules: Hydrogen Bonds. Water is of great biological importance. It is the medium in which all metabolic reactions take place in cells. Between 70% to 95% of the mass of a cell is water. Water is composed of atoms of hydrogen and oxygen. One atom of oxygen combines with two atoms of hydrogen by sharing electrons ( covalent bonding)

  19. The water cycle (article)

    Most of the water on Earth does not cycle—move from one place to another—very rapidly. We can see this in the figure below, which shows the average time that an individual water molecule spends in each of Earth's major water reservoirs, a measurement called residence time.Water in oceans, underground, and in the form of ice tends to cycle very slowly.

  20. Water Molecule

    The bonds between water molecules are the covalent bonds ;these bonds are so strong that water resists changes in its state. For example, solid, liquid and gas. Thus water has a higher melting point and a higher boiling point than any other molecule of similar size (Kirk 256). When heat is applied to solid Humans Soon to be Extinct...

  21. ScienceAdviser: There's water on Mars—but we'll probably ...

    Those models suggested that outside of the planet's frozen polar regions, the Martian crust is essentially devoid of water in any form—including ice—for about 5 kilometers. But there likely is a one- to two-kilometer-thick layer of rock rich in liquid water some 11.5 to 20 kilometers deep.