PhET Home Page

  • Sign in / Register
  • Administration
  • Edit profile

visual representation for electronegativity

The PhET website does not support your browser. We recommend using the latest version of Chrome, Firefox, Safari, or Edge.

What Is Electronegativity and How Does It Work?

ThoughtCo/Todd Helmenstine

  • Chemical Laws
  • Periodic Table
  • Projects & Experiments
  • Scientific Method
  • Biochemistry
  • Physical Chemistry
  • Medical Chemistry
  • Chemistry In Everyday Life
  • Famous Chemists
  • Activities for Kids
  • Abbreviations & Acronyms
  • Weather & Climate
  • Ph.D., Biomedical Sciences, University of Tennessee at Knoxville
  • B.A., Physics and Mathematics, Hastings College

Electronegativity is the property of an atom which increases with its tendency to attract the electrons of a bond. If two bonded atoms have the same electronegativity values as each other, they share electrons equally in a covalent bond. Usually, the electrons in a chemical bond are more attracted to one atom (the more electronegative one) than to the other. This results in a polar covalent bond. If the electronegativity values are very different, the electrons aren't shared at all. One atom essentially takes the bond electrons from the other atom, forming an ionic bond.

Key Takeaways: Electronegativity

  • Electronegativity is an atom's tendency to attract electrons to itself in a chemical bond.
  • The most electronegative element is fluorine. The least electronegative or most electropositive element is francium.
  • The greater the difference between atom electronegativity values, the more polar the chemical bond formed between them.

Avogadro and other chemists studied electronegativity before it was formally named by Jöns Jacob Berzelius in 1811. In 1932, Linus Pauling proposed an electronegativity scale based on bond energies. Electronegativity values on the Pauling scale are dimensionless numbers that run from about 0.7 to 3.98. The Pauling scale values are relative to the electronegativity of hydrogen (2.20). While the Pauling scale is most often used, other scales include the Mulliken scale, Allred-Rochow scale, Allen scale, and Sanderson scale.

Electronegativity is a property of an atom within a molecule, rather than an inherent property of an atom by itself. Thus, electronegativity actually varies depending on an atom's environment. However, most of the time an atom displays similar behavior in different situations. Factors that affect electronegativity include the nuclear charge and the number and location of electrons in an atom.

Electronegativity Example

The chlorine atom has a higher electronegativity than the hydrogen atom, so the bonding electrons will be closer to the Cl than to the H in the HCl molecule.

In the O 2 molecule, both atoms have the same electronegativity. The electrons in the covalent bond are shared equally between the two oxygen atoms.

Most and Least Electronegative Elements

The most electronegative element on the periodic table is fluorine (3.98). The least electronegative element is cesium (0.79). The opposite of electronegativity is electropositivity, so you could simply say cesium is the most electropositive element. Note that older texts list both francium and cesium as least electronegative at 0.7, but the value for cesium was experimentally revised to the 0.79 value. There is no experimental data for francium, but its ionization energy is higher than that of cesium, so it is expected that francium is slightly more electronegative.

Electronegativity as a Periodic Table Trend

Like electron affinity, atomic/ionic radius, and ionization energy, electronegativity shows a definite trend on the periodic table .

  • Electronegativity generally increases moving from left to right across a period. The noble gases tend to be exceptions to this trend.
  • Electronegativity generally decreases moving down a periodic table group. This correlates with the increased distance between the nucleus and the valence electron.

Electronegativity and ionization energy follow the same periodic table trend. Elements that have low ionization energies tend to have low electronegativities. The nuclei of these atoms don't exert a strong pull on electrons . Similarly, elements that have high ionization energies tend to have high electronegativity values. The atomic nucleus exerts a strong pull on electrons.

Jensen, William B. "Electronegativity from Avogadro to Pauling: Part 1: Origins of the Electronegativity Concept." 1996, 73, 1. 11, J. Chem. Educ., ACS Publications, January 1, 1996.

Greenwood, N. N. "Chemistry of the Elements." A. Earnshaw, (1984). 2nd Edition, Butterworth-Heinemann, December 9, 1997.

Pauling, Linus. "The Nature of the Chemical Bond. IV. The Energy of Single Bonds and the Relative Electronegativity of Atoms". 1932, 54, 9, 3570-3582, J. Am. Chem. Soc., ACS Publications, September 1, 1932.

Pauling, Linus. "The Nature of the Chemical Bond and the Structure of Molecules and Crystals: An Introduction to Mode." 3rd Edition, Cornell University Press, January 31, 1960.

  • Examples of Polar and Nonpolar Molecules
  • How to Use a Periodic Table of Elements
  • Periodic Table Definition in Chemistry
  • Printable Periodic Tables (PDF)
  • Electronegativity and Chemical Bonding
  • What Is the Most Electronegative Element?
  • Learn Which Element Has the Lowest Electronegativity Value
  • The Periodic Properties of the Elements
  • Most Reactive Metal on the Periodic Table
  • Cool Chemical Element Facts
  • Why Do Atoms Create Chemical Bonds?
  • Polar Bond Definition and Examples
  • Periodic Law Definition in Chemistry
  • The Main Types of Chemical Bonds
  • Ionic vs. Covalent Bonds: How Are They Different?
  • Chart of Periodic Table Trends

Encyclopedia Britannica

  • Games & Quizzes
  • History & Society
  • Science & Tech
  • Biographies
  • Animals & Nature
  • Geography & Travel
  • Arts & Culture
  • 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

electronegativity values of the elements in the periodic table

  • Why does physics work in SI units?
  • Is mathematics a physical science?

Highway Night Traffic Portland, drive, driving, car, automobile.

electronegativity

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

  • University of Wisconsin Pressbooks - Electronegativity
  • Chemguide - Electronegativity
  • National Center for Biotechnology Information - PubMed Central - Electronegativity determination of individual surface atoms by atomic force microscopy
  • University of Central Florida Pressbooks - Electronegativity and Polarity
  • Angelo State University - Periodic Trends — Electronegativity
  • Chemistry LibreTexts - Electronegativity
  • MHCC Library Press - Chemistry of Food and Cooking - Electronegativity and Bond Polarity
  • UEN Digital Press with Pressbooks - Electronegativity
  • Nature - Thermochemical electronegativities of the elements
  • Khan Academy - Electronegativity

electronegativity , in chemistry , the ability of an atom to attract to itself an electron pair shared with another atom in a chemical bond.

The commonly used measure of the electronegativities of chemical elements is the electronegativity scale derived by Linus Pauling in 1932. In it the elements are tabulated in decreasing order of electronegativity, fluorine being the most electronegative and cesium the least. The scale was derived from a comparison of the energies associated with chemical bonds between various combinations of atoms. A scale very similar to Pauling’s values has been obtained by measurements of atomic ionization potentials and electron affinities .

crystal bonding

Elements differing greatly in electronegativity tend to form ionic compounds , composed of positively and negatively charged units called ions; those differing moderately in electronegativity form polar, covalent compounds, in which atoms are held together by chemical bonds but which show some degree of ionization, while those elements with approximately equal electronegativities form nonpolar compounds, which show little charge separation.

Electronegativity—a perspective

  • Original Paper
  • Published: 23 July 2018
  • Volume 24 , article number  214 , ( 2018 )

Cite this article

visual representation for electronegativity

  • Peter Politzer 1 &
  • Jane S. Murray 1  

1553 Accesses

41 Citations

Explore all metrics

Electronegativity is a very useful concept but it is not a physical observable; it cannot be determined experimentally. Most practicing chemists view it as the electron-attracting power of an atom in a molecule. Various formulations of electronegativity have been proposed on this basis, and predictions made using different formulations generally agree reasonably well with each other and with chemical experience. A quite different approach, loosely linked to density functional theory, is based on a ground-state free atom or molecule, and equates electronegativity to the negative of an electronic chemical potential. A problem that is encountered with this approach is the differentiation of a noncontinuous function. We show that this approach leads to some results that are not chemically valid. A formulation of atomic electronegativity that does prove to be effective is to express it as the average local ionization energy on an outer contour of the atom’s electronic density.

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

Access this article

Price includes VAT (Russian Federation)

Instant access to the full article PDF.

Rent this article via DeepDyve

Institutional subscriptions

visual representation for electronegativity

Similar content being viewed by others

Evaluating significance in linear mixed-effects models in r.

visual representation for electronegativity

Centrality measures in networks

The bayesian new statistics: hypothesis testing, estimation, meta-analysis, and power analysis from a bayesian perspective.

Iczkowski RP, Margrave JL (1961) J Am Chem Soc 83:3547–3551

Pauling L (1932) J Am Chem Soc 54:3570–3582

Pauling L (1960) The nature of the chemical bond, 3rd edn. Cornell University Press, Ithaca

Allred AL (1961) J Inorg Nucl Chem 17:215–221

Cárdenas C, Heidar-Zadeh F, Ayers PW (2016) Phys Chem Chem Phys 18:25721–25734

Politzer P, Shields ZP-I, Bulat FA, Murray JS (2011) J Chem Theory Comput 7:377–384

Pearson RG (1968) Chem Commun 65–67

Murphy LR, Meek TL, Allred AL, Allen LC (2000) J Phys Chem A 104:5867–5871

Mulliken RS (1934) J Chem Phys 2:782–793

Mulliken RS (1935) J Chem Phys 3:573–585

Hinze J, Jaffé HH (1962) J Am Chem Soc 84:540–546

Gordy W (1946) Phys Rev 69:604–607

Allred AL, Rochow EO (1958) J Inorg Nucl Chem 5:264–268

Mande C, Deshmukh P (1977) J Phys B 10:2293–2300

Boyd RJ, Markus GE (1981) J Chem Phys 75:5385–5388

Luo Y-R, Benson SW (1992) Acc Chem Res 25:375–381

Politzer P, Grice ME, Murray JS (2001) J Mol Struct (THEOCHEM) 549:69–76

Li K, Xue D (2006) J Phys Chem A 110:11332–11337

Sanderson RT (1952) J Am Chem Soc 74:272–274

Sanderson RT (1955) Science 121:207–208

Gyftopoulos EP, Hatsopoulos GN (1968) Proc Natl Acad Sci USA 60:786–793

Hohenberg P, Kohn W (1964) Phys Rev B 136:864–871

Parr RG, Donnelly RA, Levy M, Palke WE (1978) J Chem Phys 68:3801–3807

Einhorn ME, Blankenbecler R (1971) Ann Phys 67:480

March NH (1993) Struct Bond 80:71–86

Nguyen-Dang TT, Bader RFW, Essén H (1982) Int J Quantum Chem 22:1049–1058

Hinze J (1999) Chapter 7. In: Maksic ZB, Orville-Thomas WJ (eds) Pauling’s legacy: modern modelling of the chemical bond. Elsevier, Amsterdam, pp 189–212

Chermette H (1999) J Comput Chem 20:129–154

Gopinathan MS, Whitehead MA (1980) Israel J Chem 19:209–214

Perdew JP, Parr RG, Levy M, Balduz Jr JL (1982) Phys Rev Lett 49:1691–1694

Kohn W, Becke AD, Parr RG (1996) J Phys Chem 100:12974–12980

Zhang Y, Yang W (2000) Theor Chem Accounts 103:346–348

Yang W, Zhang Y, Ayers PW (2000) Phys Rev Lett 84:5172–5175

Geerlings P, De Proft F, Langenaeker W (2003) Chem Rev 103:1793–1874

Miranda-Quintana RA, Ayers PW (2016) J Chem Phys 144:244112

Heidar-Zadeh F, Miranda-Quintana RA, Verstraelen T, Bultinck P, Ayers PW (2016) J Chem Theory Comput 12:5777–5787

Gazquez JL, Ortiz E (1984) J Chem Phys 81:2741–2748

Politzer P, Huheey JE, Murray JS, Grodzicki M (1992) J Mol Struct (THEOCHEM) 259:99–120

Politzer P, Murray JS (2006) Chem Phys Lett 431:195–198

Politzer P, Murray JS, Concha MC, Jin P (2007) Collect Czechoslov Chem Commun 72:51–63

Pearson RG (1988) Inorg Chem 27:734–740

Gázquez JL, Martínez A, Méndez F (1993) J Phys Chem 97:4059–4063

Pal S, Chandra AK, Roy RK (1996) J Mol Struct (THEOCHEM) 361:57–61

Torrent-Sucarrat M, Luis JM, Duran M, Solà M (2001) J Am Chem Soc 123:7951–7952

Zhan C-G, Nichols JA, Dixon DA (2003) J Phys Chem A 107:4184–4195

Toro-Labbé A (ed) (2007) Chemical reactivity. Elsevier, Amsterdam

Chattaraj P (ed) (2008) Theory of chemical reactivity. Taylor & Francis, Boca Raton

Liebman JF, Huheey JE (1987) Phys Rev D 36:1559–1561

Bratsch SG (1988) J Chem Ed 65:34–41

Bratsch SG (1988) J Chem Educ 65:223–227

Allen LC (1989) J Am Chem Soc 111:9003–9014

von Szentpály L (2018) J Phys Chem A 119:1715–1722

Anslyn EV, Dougherty DA (2006) Modern physical organic chemistry. University Science Books, Herndon

Parr RG, Pearson RG (1983) J Am Chem Soc 105:7512–7516

Pearson RG (1990) Acc Chem Res 23:1–2

Komorowski L (1983) Chem Phys Lett 103:201–204

Allen LC (1990) Acc Chem Res 23:175–176

Datta D, Shee NK, von Szentpály L (2013) J Phys Chem A 117:200–206

Donnelly RA, Parr RG (1978) J Chem Phys 69:4431–4439

Bergmann D, Hinze J (1996) Angew Chem Int Ed Eng 35:150–163

Sjoberg P, Brinck T, Murray JS, Politzer P (1990) Can J Chem 68:1440–1443

Politzer P, Murray JS (2007) Chapter 8. In: Toro-Labbé A (ed) Chemical reactivity. Elsevier, Amsterdam, pp 119–137

Politzer P, Murray JS, Bulat FA (2010) J Mol Model 16:1731–1742

Clark T (2010) J Mol Model 16:1231–1238

Brinck T, Carlqvist P, Stenlid JH (2016) J Phys Chem A 120:10023–10032

Stenlid JH, Brinck T (2017) J Organomet Chem 82:3072–3083

Sacher E, Currie JF (1988) J Electron Spectrosc Relat Phenom 46:173–177

Luo Y-R, Benson SW (1989) J Phys Chem 93:7333–7335

DeKock RL (1990) J Phys Chem 94:1713–1714

Allen LC (1994) Int J Quantum Chem 49:253–277

Allen LC (1998) In: PvR S (ed) Encyclopedia of computational chemistry, vol 2. Wiley, New York, pp 835–852

Mann JB, Meek TL, Allen LC (2000) J Am Chem Soc 122:2780–2783

Mann JB, MeekTL KET, Capitani JF, Allen LC (2000) J Am Chem Soc 122:5132–5137

Jorgensen CK (1971) Chimia 25:213

Politzer P, Daiker KC (1973) Chem Phys Lett 20:309–316

Huheey JE, Keiter EA, Keiter RL (1993) Inorganic chemistry: principles of structure and reactivity, 4th edn. HarperCollins, New York

Politzer P, Murray JS, Grice ME (2005) Collect Czechoslov Chem Commun 70:550–558

Koopmans TA (1934) Physica 1:104–113

Nesbet RK (1965) Adv Chem Phys 9:321–363

Bader RFW, Carroll MT, Cheeseman JR, Chang C (1987) J Am Chem Soc 109:7968–7979

Politzer P, Abu-Awwad F, Murray JS (1998) Int J Quantum Chem 69:607–613

Politzer P, Murray JS, Concha MC (2002) Int J Quantum Chem 88:19–27

Ryabinkin IG, Staroverov VN (2014) J Chem Phys 141:084107 1–8

Kohut SV, Cuevas-Saavedra R, Staroverov VN (2016) J Chem Phys 145:074113

Delgado-Barrio G, Prat RF (1975) Phys Rev A 12:2288–2297

Politzer P, Murray JS, Grice ME, Brinck T, Ranganathan S (1991) J Chem Phys 95:6699–6704

Murray JS, Politzer P (2009) Croat Chim Acta 82:267–275

Perdew JP, Burke K, Ernzerhof M (1996) Phys Rev Lett 77:3865–3868

Lee C, Yang W, Parr RG (1988) Phys Rev B 37:785–789

Becke AD (1993) J Chem Phys 98:5648–5652

Zhao Y, Truhlar DG (2008) Theor Chem Accounts 120:215–241

Zhao Y, Truhlar DG (2008) Acc Chem Res 41:157–167

Alonso JA, Girifalco LA (1980) J Chem Phys 73:1313–1319

March NH, Bader RFW (1980) Phys Lett 78A:242–243

Tkacz-Śmiech K, Ptak WS, Koleżyński A, Mrugalski J (1994) Int J Quantum Chem 51:569–575

Pearson RG (1992) Inorg Chim Acta 198-200:781–786

Download references

Author information

Authors and affiliations.

Department of Chemistry, University of New Orleans, New Orleans, LA, 70148, USA

Peter Politzer & Jane S. Murray

You can also search for this author in PubMed   Google Scholar

Corresponding author

Correspondence to Peter Politzer .

Rights and permissions

Reprints and permissions

About this article

Politzer, P., Murray, J.S. Electronegativity—a perspective. J Mol Model 24 , 214 (2018). https://doi.org/10.1007/s00894-018-3740-6

Download citation

Received : 10 February 2018

Accepted : 27 June 2018

Published : 23 July 2018

DOI : https://doi.org/10.1007/s00894-018-3740-6

Share this article

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

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

Provided by the Springer Nature SharedIt content-sharing initiative

  • Electronegativity
  • Electronic chemical potential
  • Average local ionization energy
  • Find a journal
  • Publish with us
  • Track your research

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

  • View all journals
  • Explore content
  • About the journal
  • Publish with us
  • Sign up for alerts
  • Open access
  • Published: 26 April 2017

Electronegativity determination of individual surface atoms by atomic force microscopy

  • Jo Onoda 1 , 2 ,
  • Martin Ondráček 3 ,
  • Pavel Jelínek 3 , 4 &
  • Yoshiaki Sugimoto 1 , 2  

Nature Communications volume  8 , Article number:  15155 ( 2017 ) Cite this article

12k Accesses

41 Citations

44 Altmetric

Metrics details

Atomic force microscopy

  • Characterization and analytical techniques
  • Scanning probe microscopy

Electronegativity is a fundamental concept in chemistry. Despite its importance, the experimental determination has been limited only to ensemble-averaged techniques. Here, we report a methodology to evaluate the electronegativity of individual surface atoms by atomic force microscopy. By measuring bond energies on the surface atoms using different tips, we find characteristic linear relations between the bond energies of different chemical species. We show that the linear relation can be rationalized by Pauling’s equation for polar covalent bonds. This opens the possibility to characterize the electronegativity of individual surface atoms. Moreover, we demonstrate that the method is sensitive to variation of the electronegativity of given atomic species on a surface due to different chemical environments. Our findings open up ways of analysing surface chemical reactivity at the atomic scale.

Similar content being viewed by others

visual representation for electronegativity

Imaging surface structure and premelting of ice Ih with atomic resolution

visual representation for electronegativity

Label-free detection and profiling of individual solution-phase molecules

visual representation for electronegativity

Understanding how junction resistances impact the conduction mechanism in nano-networks

Introduction.

Electronegativity, an important theoretical concept in chemistry, was originally defined by Linus Pauling as ‘the power of an atom in a molecule to attract electrons to itself’ 1 , 2 . Since that time, various classical scales of electronegativity have been suggested 3 , 4 , 5 , 6 , including the rigorous modern formalism of absolute electronegativity developed along with density functional theory (DFT) 7 . In contrast, conventional experimental means for measuring electronegativity have been limited to thermochemical techniques, that is, the measurement of ensemble-averaged bond energies 8 . On the other hand, atomic force microscopy (AFM) was able to achieve atomic resolution on both semiconductor 9 and insulator 10 surfaces. It has also found numerous applications in the field of chemistry: for example, quantitative measurement of short-range chemical forces 11 , chemical identification of individual surface atoms 12 , visualization of the internal structures of organic molecules 13 , 14 , 15 , 16 and metal clusters 17 , discrimination of the Pauling bond order 18 , 19 and tracking of surface chemical reactions 20 , 21 .

Here, we extend this already impressive suite of applications by demonstrating the use of AFM in characterizing the electronegativity of individual surface atoms. Pauling’s electronegativity values for individual single atoms on surfaces can be estimated using site-specific energy spectroscopy acquired with a variety of AFM tips. Namely, we show that the binding energies for individual surface atoms observed using different tips can provide an ensemble of data that can be used later to determine the electronegativity of different chemical elements or chemical groups on a surface ( Fig. 1a ). Our experimental findings are supported by theoretical analysis based on DFT calculations.

figure 1

( a ) Schematic illustration of AFM energy spectroscopy with the polar covalent bond of Si–O. ( b ) Δ f curves measured on Si and O adatoms on the Si(111)-(7 × 7) surface. The background (BG) Δ f ( z ) curve is also included for subtraction of the long-range component (Methods). ( c ) E ( z ) curves converted from short-range components in the Δ f ( z ) curves in b . Bond energies are defined at the minima of E ( z ) curves. The acquisition parameters are f 0 =152.913 kHz, the oscillation amplitude ( A )=146 Å, the spring constant ( k )=28.2 N m −1 and sample bias ( V S )=+40 mV.

Scatter plots of the bond energies

figure 2

Scatter plots of the bond energies of Ge ( a ), Sn ( b ), Al ( c ), O ( d ) adatoms, SiO 2 ( e ) and SiNO ( f ) complexes obtained experimentally. The bond energies were measured above the bright spots in AFM images as shown in the insets. The individual error bars are estimated as 10% of the corresponding bond energies based on uncertainties in measurements of A and k . The scatter plots were fitted using weighted orthogonal distance regression.

Interpretation of the linear relations by Pauling’s equation

We rationalized the linear relation between the bond energy of tip apexes and different surface atoms by employing Pauling’s equation for the bond energy E A–B of heterogeneous polar bonds between atoms A and B 1 , 2 :

Chemical identification by the slopes

Electronegativity determination by the intercepts.

figure 3

( a ) A one-dimensional plot showing the variation of the Hartree potential above different adatoms such as Si, Ge, Al, O or N. The zero point of the Hartree potential is set to the Fermi level, and the z -distance is aligned to the height of an inspected adatom ( z X ) on the Si(111)-(7 × 7) surface. The calculated electron density isosurfaces (0.02 e Å −3 in yellow) of the highest occupied frontiers orbital on the Si adatom ( b ), SiO 2 ( c ) around the Fermi level and SiNO ( d ) located 1 eV below the Fermi level. Cut-plane of the electron density isosurface is coloured from red (0.13 e Å −3 ) to blue (0.02 e Å −3 ). Plotted two-dimensional plane of the Hartree potential projected onto the Si adatom ( e ), SiO 2 ( f ) and SiNO ( g ) complex (range from 0 (blue) to 5 (red) eV).

Evaluation of group electronegativity

In general, the electronegativity of surface atoms (here, Si adatoms) can be modified by their surrounding environments: for example, structural and chemical rearrangement of local structures, charge transfer from the neighbouring atoms and re-hybridization of orbitals. Previously, this information has not been experimentally accessible. From this perspective, the present method provides a unique opportunity to obtain electronegativity values for specific surface atoms in different local chemical structures. To demonstrate such sensitivity towards group electronegativity, we carried out measurements on a ‘SiO 2 ’ structure ( Fig. 2e ), where two O atoms are inserted into the back bonds of an Si adatom 22 , and on a ‘SiNO’ structure ( Fig. 2f ) created from dissociated O and N atoms ( Supplementary Methods 1 ). Local atomic arrangements obtained from the total-energy DFT calculations of the SiO 2 and SiNO structures are presented in Fig. 3c,d , respectively.

In conclusion, we have presented here a method that, for the first time to our knowledge, allows experimental characterization of the electronegativity of individual surface atoms by means of AFM. We have found that the scatter plots of the bond energies can be interpreted based on Pauling’s equation for polar covalent bonds. This allowed us to disentangle the covalent and ionic bond energies in polar covalent bonds, and hence to estimate the electronegativity of the individual elements. In addition, this method facilitates analysis of the group electronegativity of atoms belonging to locally formed nanostructures. We expect that the present method can be applied to other interesting systems such as transition metal oxides used in catalysis by terminating AFM tips with known atoms on the surface. Furthermore, we believe that the method can not only characterize the specific electronegativity of individual atoms, defects, adsorbates and dopant impurities on surfaces but may also open up new ways of analysing surface chemical reactivity in terms of surface electrophilicity/nucleophilicity against reactants 29 , and surface chemical softness/hardness with regard to the hard–soft acid–base principle 26 , 30 .

All experiments were carried out using a custom-built ultrahigh-vacuum AFM system at room temperature with a typical base pressure of 5 × 10 −9  Pa. An optical interferometer was equipped to detect the cantilever deflection. Frequency-modulation technique was used for obtaining Δ f . We used commercial Si cantilevers, whose tip apexes were cleaned by Ar ion sputtering to remove the native oxidation layers. The cantilever was oscillated with sufficiently large amplitude ( A ∼ 100 Å) for stable AFM operation. We chose the Si(111)-(7 × 7) surface as playground for the purpose of the present experiments since the surface had many Si adatoms on which atoms at tip apex were replaced by mild contact. To compensate a contact potential difference between tip and sample, we properly applied V S to the sample while keeping the tip bias at ground.

Site-specific energy spectroscopy

When we performed site-specific energy spectroscopy, thermal drift was compensated by feed-forward technique 31 and the precision of lateral position of the point spectroscopy became better than ±0.1 Å, both of which were realized by our atom-tracking implementation 32 . After recording several Δ f ( z ) curves on an adatom of interest, we averaged them to a single Δ f ( z ) curve for converting to E ( z ). Note that energy dissipation during the spectroscopy was negligible. For eliciting accurate bond energies exerted on the foremost tip atom and surface adatoms, we subtracted the background Δ f ( z ) curve ( Fig. 1b ; Supplementary Methods 1 ), which is obtained by analytically fitting a Δ f ( z ) curve measured on a corner hole site 11 , 12 , from Δ f ( z ) curves on the individual target and reference adatom sites. Finally, the Δ f ( z ) curves only containing the short-range contribution were converted to E ( z ) using the inversion procedure suggested by Sader and Jarvis 33 .

Density functional theory calculations

All DFT calculations have been done using the Vienna Ab-Initio Simulation Package (VASP 4.6) 34 , 35 , which is a pseudopotential plane-wave code. The basic versions of ultrasoft Vanderbilt pseudopotentials 36 , 37 supplied with VASP were used for all chemical elements involved (Si, H, O, N, Al and Ge) apart from Sn, for which the Sn d potential that includes d -electrons as valence electrons was adopted. The PW91 implementation of the generalized gradient approximation 38 was chosen to describe the exchange-correlation functional. The size of the plane-wave basis was determined by the cutoff energy of 300 or 396 eV. Only the central (Gamma) point of the first Brillouin zone was considered for the Bloch-wave solutions.

Data availability

The data that support the findings of this study are available from the corresponding author on reasonable request.

Additional information

How to cite this article: Onoda, J. et al . Electronegativity determination of individual surface atoms by atomic force microscopy. Nat. Commun. 8, 15155 doi: 10.1038/ncomms15155 (2017).

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

Pauling, L. The nature of the chemical bond. IV. The energy of single bonds and the relative electronegativity of atoms. J. Am. Chem. Soc. 54 , 3570–3582 (1932).

Article   CAS   Google Scholar  

Pauling, L. in The Nature of the Chemical Bond 3rd, 13th Printing 1995 edn (Cornell University Press, 1960).

Mulliken, R. S. A new electroaffinity scale; together with data on valence states and on valence ionization potentials and electron affinities. J. Chem. Phys. 2 , 782–793 (1934).

Article   CAS   ADS   Google Scholar  

Sanderson, R. T. Electronegativities in inorganic chemistry. J. Chem. Educ. 29 , 539–544 (1952).

Allred, A. L. & Rochow, E. G. A scale of electronegativity based on electrostatic force. J. Inorg. Nucl. Chem. 5 , 264–268 (1958).

Gordy, W. A new method of determining electronegativity from other atomic properties. Phys. Rev. 69 , 604–607 (1946).

Parr, R. G., Donnelly, R. A., Levy, M. & Palke, W. E. Electronegativity: the density functional viewpoint. J. Chem. Phys. 68 , 3801–3807 (1978).

Wagman, D. D. et al. in The NBS Tables of Chemical Thermodynamic Properties: Selected Values for Inorganic and C1 and C2 Organic Substances in SI Units (American Chemical Society and American Institute of Physics, 1982).

Giessibl, F. J. Atomic resolution of the silicon (111)-(7 × 7) surface by atomic force microscopy. Science 267 , 68 (1995).

Barth, C. & Reichling, M. Imaging the atomic arrangements on the high-temperature reconstructed α-Al2O3(0001) surface. Nature 414 , 54–57 (2001).

Lantz, M. A. et al. Quantitative measurement of short-range chemical bonding forces. Science 291 , 2580–2583 (2001).

Sugimoto, Y. et al. Chemical identification of individual surface atoms by atomic force microscopy. Nature 446 , 64–67 (2007).

Gross, L., Mohn, F., Moll, N., Liljeroth, P. & Meyer, G. The chemical structure of a molecule resolved by atomic force microscopy. Science 325 , 1110–1114 (2009).

Gross, L. et al. Organic structure determination using atomic-resolution scanning probe microscopy. Nat. Chem. 2 , 821–825 (2010).

Mohn, F., Gross, L., Moll, N. & Meyer, G. Imaging the charge distribution within a single molecule. Nat. Nanotechnol. 7 , 227–231 (2012).

Iwata, K. et al. Chemical structure imaging of a single molecule by atomic force microscopy at room temperature. Nat. Commun. 6 , 7766 (2015).

Emmrich, M. et al. Subatomic resolution force microscopy reveals internal structure and adsorption sites of small iron clusters. Science 348 , 308–311 (2015).

Gross, L. et al. Bond-order discrimination by atomic force microscopy. Science 337 , 1326–1329 (2012).

Pavliček, N. et al. On-surface generation and imaging of arynes by atomic force microscopy. Nat. Chem. 7 , 623–628 (2015).

Article   Google Scholar  

de Oteyza, D. G. et al. Direct imaging of covalent bond structure in single-molecule chemical reactions. Science 340 , 1434–1437 (2013).

Schuler, B. et al. Reversible bergman cyclization by atomic force microscopy. Nat. Chem. 8 , 220–224 (2016).

Onoda, J., Ondráček, M., Yurtsever, A., Jelínek, P. & Sugimoto, Y. Initial and secondary oxidation products on the Si(111)-(7 × 7) surface identified by atomic force microscopy and first principle calculations. Appl. Phys. Lett. 104 , 133107 (2014).

Article   ADS   Google Scholar  

Setvín, M. et al. Chemical identification of single atoms in heterogeneous III-IV chains on Si(100) surface by means of nc-AFM and DFT calculations. ACS Nano 6 , 6969–6976 (2012).

Onoda, J., Niki, K. & Sugimoto, Y. Identification of Si and Ge atoms by atomic force microscopy. Phys. Rev. B 92 , 155309 (2015).

Pou, P. et al. Structure and stability of semiconductor tip apexes for atomic force microscopy. Nanotechnology 20 , 264015 (2009).

Parr, R. G. & Pearson, R. G. Absolute hardness: companion parameter to absolute electronegativity. J. Am. Chem. Soc. 105 , 7512–7516 (1983).

Sadewasser, S. & Glatzel, T. Kelvin Probe Force Microscopy (Springer-Verlag (2012).

Sadewasser, S. et al. New insights on atomic-resolution frequency-modulation Kelvin-probe force-microscopy imaging of semiconductors. Phys. Rev. Lett. 103 , 266103 (2009).

Brommer, K. D., Galván, M., Pino, A. D. Jr & Joannopoulos, J. D. Theory of adsorption of atoms and molecules on Si(111)-(7 × 7). Surf. Sci. 314 , 57–70 (1994).

Pearson, R. G. Hard and soft acids and bases. J. Am. Chem. Soc. 85 , 3533–3539 (1963).

Abe, M. et al. Drift-compensated data acquisition performed at room temperature with frequency modulation atomic force microscopy. Appl. Phys. Lett. 90 , 203103 (2007).

Abe, M., Sugimoto, Y., Custance, O. & Morita, S. Room-temperature reproducible spatial force spectroscopy using atom-tracking technique. Appl. Phys. Lett. 87 , 173503 (2005).

Sader, J. E. & Jarvis, S. P. Accurate formulas for interaction force and energy in frequency modulation force spectroscopy. Appl. Phys. Lett. 84 , 1801–1803 (2004).

Kresse, G. & Furthmüller, J. Efficiency of ab initio total energy calculations for metals and semiconductors using a plane-wave basis set. Comput. Mater. Sci. 6 , 15–50 (1996).

Kresse, G. & Furthmüller, J. Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys. Rev. B 54 , 11169–11186 (1996).

Vanderbilt, D. Soft self-consistent pseudopotentials in a generalized eigenvalue formalism. Phys. Rev. B 41 , 7892–7895 (1990).

Kresse, G. & Hafner, J. Norm-conserving and ultrasoft pseudopotentials for first-row and transition-elements. J. Phys.: Condens. Matter 6 , 8245–8257 (1994).

CAS   ADS   Google Scholar  

Perdew, J. P. et al. Atoms, molecules, solids, and surfaces: applications of the generalized gradient approximation for exchange and correlation. Phys. Rev. B 46 , 6671–6687 (1992).

Download references

Acknowledgements

This work was supported by Grants-in-Aid for JSPS Fellows (14J00689) and for Scientific Research (15H03566, 16H00933, 16H00959, 16K13680, 16K17482) from the Ministry of Education, Culture, Sports, Science, and Technology of Japan (MEXT). We also acknowledge financial support from the Asahi Glass Foundation, the Mitsubishi Foundation, the Yamada Science Foundation, Ube foundation, and the Precise Measurement Technology Promotion Foundation. P.J. and M.O. acknowledge financial support of GACR (Czech Science Foundation) under Project no. 14-16963J and Czech Academy of Sciences through Praemium Academiae award.

Author information

Authors and affiliations.

Department of Advanced Materials Science, Graduate School of Frontier Sciences, University of Tokyo, 5-1-5 Kashiwanoha, Kashiwa, 277-8561, Chiba, Japan

Jo Onoda & Yoshiaki Sugimoto

Division of Electrical, Electronic and Information Engineering, Graduate School of Engineering, Osaka University, 2-1 Yamada-Oka, Suita, 565-0871, Osaka, Japan

Department of Thin Films and Nanostructures, Institute of Physics, Czech Academy of Sciences, Cukrovarnická 10/112, Prague, 16200, Czech Republic

Martin Ondráček & Pavel Jelínek

Department of Physical Chemistry, Regional Centre of Advanced Technologies and Materials, Palacky University, Šlechtitelů 27, Olomouc, 78371, Czech Republic

Pavel Jelínek

You can also search for this author in PubMed   Google Scholar

Contributions

J.O. and Y.S. conceived and designed the experiments. J.O. and Y.S. performed the experiments. M.O. and P.J. performed the theoretical analysis. J.O. and P.J. wrote the paper. All authors discussed the results and commented on the manuscript.

Corresponding author

Correspondence to Jo Onoda .

Ethics declarations

Competing interests.

The authors declare no competing financial interests.

Supplementary information

Supplementary information.

Supplementary Figures, Supplementary Notes, Supplementary Methods and Supplementary References (PDF 1368 kb)

Rights and permissions

This work is licensed under a Creative Commons Attribution 4.0 International License. The images or other third party material in this article are included in the article’s Creative Commons license, unless indicated otherwise in the credit line; if the material is not included under the Creative Commons license, users will need to obtain permission from the license holder to reproduce the material. To view a copy of this license, visit http://creativecommons.org/licenses/by/4.0/

Reprints and permissions

About this article

Cite this article.

Onoda, J., Ondráček, M., Jelínek, P. et al. Electronegativity determination of individual surface atoms by atomic force microscopy. Nat Commun 8 , 15155 (2017). https://doi.org/10.1038/ncomms15155

Download citation

Received : 14 November 2016

Accepted : 27 February 2017

Published : 26 April 2017

DOI : https://doi.org/10.1038/ncomms15155

Share this article

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

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

Provided by the Springer Nature SharedIt content-sharing initiative

This article is cited by

Observation of electron orbital signatures of single atoms within metal-phthalocyanines using atomic force microscopy.

  • Pengcheng Chen
  • Dingxin Fan

Nature Communications (2023)

Direct assessment of the acidity of individual surface hydroxyls

  • Margareta Wagner
  • Bernd Meyer
  • Ulrike Diebold

Nature (2021)

By submitting a comment you agree to abide by our Terms and Community Guidelines . If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.

Quick links

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

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

visual representation for electronegativity

WebElements logo

  • Full property list
  • Atomic numbers
  • Atomic radii
  • Atomic spectra
  • Atomic weights
  • Block in periodic table
  • Boiling point
  • Bond enthalpies
  • Bond length in element
  • Bulk modulus
  • CAS Registry ID
  • Classification
  • Covalent radii
  • Critical temperature
  • Crystal structure
  • Density of solid
  • Description
  • Effective nucl. charges
  • Electron bind. energies
  • Electrical resistivity
  • Electron affinity
  • Electron shell structure
  • Electronegativity
  • Electronic configuration
  • Enthalpy: atomization
  • Enthalpy of fusion
  • Expansion coefficient
  • Enthalpy: vaporization
  • Group numbers
  • Health hazards
  • History of the element
  • Hydration enthalpies
  • Hydrolysis constants
  • Ionization energies
  • Lattice energies
  • Liquid Range
  • Mass in human
  • Meaning of name
  • Melting points
  • Molar volume
  • Names and symbols
  • Orbital_properties
  • Oxidation numbers
  • Period numbers
  • Poisson's ratio
  • Ionic radii
  • Radii - ionic, Pauling
  • Radii - metallic (12)
  • Radii - valence orbital
  • Radii - Van der Waals
  • Reactions of elements
  • Reduction potentials
  • Reflectivity
  • Refractive index
  • Rigidity modulus
  • Standard state
  • Superconductivity
  • Term symbol
  • Thermal conductivity
  • Thermodynamics
  • Velocity of sound
  • X-ray mass abs. coeff.
  • Young's modulus
  • Periodicity --> -->