Abstract
The chemical bond is the biggest paradigm in chemistry. But at the same time it is notoriously complex to define it. Under the assumption that knowing what we do not know makes better students and lecturers, we succinctly describe three approaches to define the bond (energetic, structural, and orbital), their advantages and especially their failures. We propose that these definitions, including their theoretical, practical and even philosophical issues, should be taught to advanced undergraduate chemistry students as an essential introduction to the chemical bond module of quantum chemistry courses. This is a controversial topic due to the lack of consensus in the chemical bond community over this heavily disputed topic and the conflicting pedagogical approach; however, and maybe because of this controversy, the teaching of the definitions of the chemical bond including their open questions and challenges can be positively instructive. Moreover, we propose that teaching the uncertainties of chemistry, especially in relation to the paradigmatic chemical bond, is an ethical matter.
1 Introduction
Theoretically, we can solve the Schrödinger equation and voilà, we have all the information of a molecule. Or, as Dirac famously put it:
The underlying physical laws necessary for the mathematical theory of a large part of physics and the whole of chemistry are thus completely known, and the difficulty is only that the exact application of these laws leads to equations much too complicated to be soluble. (Dirac, 1929)
This is what we tell chemistry students as they start college, and up to quantum chemistry courses.
Is chemistry really that simple? If we focus on the concept of molecules and their chemical bonds, indeed we would be able to obtain all their numerical properties by solving the appropriate equations; we (ideally) get the energies and other values due to Hermitian operators (like the Hamiltonian), and at the same price we may obtain the wave function. But this will not respond to conceptual notions, such as what on earth is a chemical bond. Or, as Coulson defiantly said: “I described a bond, a normal simple chemical bond; and I gave many details of its character. Sometimes it seems to me that a bond between two atoms has become so real, so tangible, so friendly that I can almost see it. And then I awake with a little shock: for a chemical bond is not a real thing, it does not exist, no-one has ever seen it, no-one ever can. It is a figment of our own imagination.” (Coulson, 1955).
Except for specific cases, in general we are missing an important side of chemical education by not teaching and debating history and philosophy of chemistry (Erduran & Kaya, 2019; Garritz, 2013). In particular, and regarding the present article, it is our belief that one important aspect that is lacking in chemical education (and in chemistry in general) is the recognition of how much we do not know about the chemical bond. There is a large and solid body of literature about teaching this topic (Dulmen et al., 2023; Hunter et al., 2022; Levy Nahum et al., 2010; Ünal et al., 2006), with most of these sources emphasizing the best strategies to teach the bond and to test the students on the topic, as well as bringing tools to minimize the students’ lacunas in knowledge, comprehension, and metacognition from high-schoolers to advanced undergraduates (Burrows & Mooring, 2015; Çalik et al., 2024; Coll & Taylor, 2001; Croft & de Berg, 2014; Hilton & Nichols, 2011; Kaldaras et al., 2024; Lee & Cheng, 2014; Legodi, 2021; Nahum et al., 2007, 2013]; Tsaparlis, 1997; Tsaparlis et al., 2018; Venkataraman, 2017; Yayon et al., 2012; Zohar & Levy, 2019). However, to the best of our knowledge, these works do not incorporate the teaching of the scientific uncertainties of the bond, that is the “known unknowns” for researchers, but “unknown unknowns” for students. And with time this is followed by a lack of introspection in the nature of the chemical bond by the general chemistry practitioners.
There is an academic myth that believes that chemical theory researchers perfectly know what binds atoms in a molecule. This is disproved by the amount of heated open discussions in the literature about what makes the bonds (Clark et al., 2022; Crabtree, 2017; Deveson et al., 2016; Frenking, 2023; Levine & Head-Gordon, 2020; Nordholm & Bacskay, 2020; Ryan, 2024; Seifert, 2022a; Sterling et al., 2024; Tarannam et al., 2021). Another fantasy shared by students and teachers alike is that when advancing from introductory to advanced courses the student will slowly improve the basic and primitive models of the bond, starting from the covalent and ionic bonds from a Lewis shared-electrons perspective, and ending in the comprehension of Schrödinger equation applied to molecules. This naïve idea creates a conviction that college fills the missing concepts step by step, and in the end, everything will be accurately explained and understood. If not even specialists know all the facets of the bond, this myth also cannot be true, and students should know it.
Such a pedagogical approach is metaphorically described by the “Neurath boat”, that is the gradual changes of taught concepts (Lewis and octet → VSEPR → MOs → resonance, hyperconjugation, etc. → Schödinger, Heisenberg, Pauli → Hückel, Hartree–Fock, DFT → …) as if we were substituting the old beams of a boat by new ones while navigating the seas (Neurath, 1973). In more mundane terms a similar analogy is the “lie to children” principle (Stewart & Cohen, 1997), where simplistic and strictly untrue [but useful! (Box, 1979)] models are used to reach the students at their level, slowly increasing the complexity of the models. It is an iterative lie, always telling the students that “in the next course you will learn the final truth about the chemical bond”, a truth that never comes. How could it, if not even specialists know all the facets of the bond, or even converge on a definition of it?
In this sense, teaching the uncertainties in the chemical bond is an ethical issue in the pedagogy of chemistry. Most topics on ethics in chemistry involve the ethical practice of this form of knowledge, such as scientific integrity and respect, sustaining life, safety in the laboratory, sustainability, ethical experimentation and publication, ethical industrial production, ethical technological development, etc. Another more modern aspect of ethics in chemistry involves behaviors common to the rest of society, including the promotion of diversity, the prevention of sexual harassment, meritocracy, or the correct interaction in imbalanced power relationships (teacher-student or master-apprentice). Consequently, chemical ethics is taught as a set of rules for the social responsibility of the future professional (Frank et al., 2011; Keinan, 2022; Kovac, 1996, 2006]; Mann, 2017; McClure & Lucius, 2010; McGuffin, 2008; Mehlich et al., 2017; Pimple, 2002; Schelble & Elkins, 2021; Schummer & Børsen, 2022; Solà, 2024; Stojanovska, 2024; Sweeting, 1999). Much rarer is the discussion of the ethical manner in which to teach chemistry, not in terms of content and syllabus, relationship with students, or testing manner, but in terms of the connection between the chemical concepts and the reality in industry and research, and the acknowledgment of their realities. As examples of this problem, we can point to the shock that students suffer when hopping from the educational to the research laboratories (in the former things must always work to learn the techniques, while in the latter things most often do not work because of unknown effects), the difficulties when passing from a passive learner to an active producer of chemical knowledge in graduate school, or the topic in question here, the failure to teach what chemists ignore.
This article was written following an advanced undergraduate course on the chemical Bond. Being a quantum chemistry class, this course approaches the bonds by describing in broad conceptual terms many types of interactions (covalent, non-covalent, ionic, dative, H-bond, halogen and other hole interactions, hyperconjugation, coordinative, metallic, van der Waals in all its forms, etc.), as well as brief explanations of techniques to compute them (Hückel, Hartree Fock, configuration of interactions, and DFT) and other topics connected with the bond (potential energy surface, reactivity, orbitals, etc.). Considering that the students belonged to all chemistry programs (organic, physical, or bio-chemistry) the syllabus relied much more on general and heuristic basic concepts and less on complex mathematical problem-solving. But the first part of the course, which is what this article is about, takes the infrequent approach of trying to define and understand what the bond is, starting with the premise that chemistry students deserve to be told the present state of affairs about the current scientific knowledge of the chemical bond, and the debates that are taking place. Even if the students already come with basic and practical knowledge of the existence and nature of the bond from previous courses, many of them receive this semi-philosophical methodology with surprise, but also with curiosity and interest. Being a short chemistry course and not a philosophy of science course, we did not go deeper into the philosophical field. However, I believe that students can gain much insight and critical thinking skills by learning the basic works of philosophers and historians of chemistry and science (Kozuch, 2022).
Based on this idea, and with the belief that knowing what we do not know can lead to better students, teachers, and researchers by improving cognition and metacognition (Lavi et al., 2019), we will try to succinctly show a small part of what we know, but especially what we do not know about how to define the chemical bond and understand its existence. We cannot and will not be extensive, since the chemical bond field deserves an encyclopaedia (Frenking & Shaik, 2014). As such, we might leave more questions than answers, which is not necessarily a bad thing. As this article discusses a highly controversial topic, numerous researchers (and reviewers) do not agree with many concepts postulated here. Again, this is not necessarily bad, as it shows the living disputes in this field which advanced chemistry students should also be aware of.
2 The three epistemological axioms
Considering the critical nature of the chemical bond concept for chemists, and the above-mentioned issues, we propose (partly as tongue-in-cheek) these “axioms” of the chemical bond, which we believe should be, in some way, taught to advanced chemistry students:
1st axiom: The chemical bond is the biggest paradigm in chemistry.
2nd axiom: No chemist is a real chemist without understanding the chemical bond.
3rd axiom: No one really understands the chemical bond.
2.1 1st axiom: the paradigm
It is a good homework to ruminate on how the existence of the chemical bond is our biggest paradigm. Indeed, the idea of the bond provides the scientific basis of chemistry and its jargon, it tells us how to carry out and interpret experiments, and it even defines what are the valid questions that we can ask when carrying out a chemical project. Or paraphrasing a famous Argentinian quote: “Within the bond everything, outside the bond nothing”.
Let us focus on the issue of what is a valid question in chemistry. For instance, when studying a chemical reaction, the fundamental (and sometimes the only) question to formulate is how the chemical bonds are reshuffled during the reaction. If we do not have chemical bonds, would that question make sense? What question would we ask if not? How is phlogiston released? How does the water:air:earth:fire ratio change?
Every question about structure, reactivity, nanochemistry, analytical chemistry, DNA, organometallics, the ozone layer, IR spectroscopy, or astrochemistry only makes sense in terms of having bonds, which gives us the language to describe chemistry. Is there a chemical field where the chemical bond is not its obvious starting point? Chemistry students are usually not aware of the weight of this chemical paradigm, and how much it affects the whole building of their studies. It is a good practice to start examining this concept from introduction to chemistry, and never stop pondering about it.
To be fair, many consider as the biggest paradigm in chemistry not the chemical bond, but the atom as the basic building block of matter. The existence of atoms, championed originally by Dalton (and maybe Democritus if we extend ourselves), applied theoretically by Boltzmann, and then finally proven by Einstein and Perrin, is at the heart of chemistry. From there it was natural to think of a “field” that bring those atoms together into molecules, even if its reason was hardly understood [it was considered a kind of blending of the “atmospheres of caloric” in the early 19th century (Grossman, 2017)]. However, in our opinion the essence of chemistry is the qualitative and quantitative way that the elements bind, while atoms “merely” are the basic components of chemistry. Or in metaphorical terms, mortar maketh the house, not bricks. In any case, the two ideas are not mutually exclusive, and we are also allowed to consider a greater chemical paradigm as “matter made by atoms joined by their chemical bonds”.
2.2 2nd axiom: the real chemist
The second epistemological axiom is about the vital importance of knowing, teaching, and learning the concept of the chemical bond. If it is the central paradigm of chemistry, it is not only important; it is inevitable. The chemical bond, with its orbitals, electrostatic potential, electron density, and energy patterns, provides the ABC of chemistry work, research, and education, and nothing can be built without it. Therefore, no chemist is a real chemist without understanding the chemical bond.
The only thing that stands for debate is how deep are we going to teach the subject (mostly in terms of how much quantum mechanics can we cover), or how broad are we going to go (how many types of bonds and effects can we cover). Undoubtedly, a better knowledge of chemical bond theory makes us better chemists, but the teaching time and the students’ ability to digest information is limited. Moreover, the approach that a lecturer chooses might depend on the target, whether it is directed to physical chemistry graduate students or if it is a general chemistry course. In our specific chemical bond course, which as explained above was directed to the general chemistry student, the focus was on the broad but relatively shallow conceptual and quantum mechanical description of as many types of bonds and effects as possible. Nevertheless, many articles, techniques and novel ideas were brought as examples, a way of showing the deepness of state-of-the-art chemical bond research and how they tackle its knowledge holes.
2.3 3rd axiom: we know nothing
The last axiom clearly is the most controversial and subjective. Simply put, it is our belief that no chemist really comprehend the chemical bond in its totality (and probably we will never will). According to a strict use of logic, of course the second and third axiom taken together imply that there are no real chemists; but needless to say, herein we are speaking figuratively.
The fact that we do not fully comprehend the bond does not mean that we know nothing about the bonds. On the contrary, we know a lot; we have a firm theoretical basis of their existence and nature based on a plethora of experimental and computational studies. For example, we know that the bond mostly is a subtle equilibrium between electrons-nuclei attractions, nuclei–nuclei and e−–e− Coulomb and Pauli repulsions, and kinetic contributions that reach a minimum in the potential energy surface. However, the point is that:
We certainly do not know everything. There is still much to learn.
We are unsure of the real nature of many chemical bonds. Some bonds are hotly debated, with some people being sure of one explanation, and others being sure of the opposite explanation.
Although the existence of the chemical bond is not debated (it is a paradigm), its definition is very far from uncontested (as we shall see below).
We know how to compute bonds, and we know the impact of each component of the Hamiltonian when creating molecules. But this gives us the “how much” information on the bond, while we still debate what are bonds in general terms.
For a surprisingly large number of intrinsic concepts of the chemical bond there are also multiple definitions that play against the clarity and unambiguity that one would expect from chemistry as an exact science (such as bond strength and energy, bond order, atomic charge, molecular orbitals, etc.).
How is it possible that we know so much about the chemical bond, and yet we know so little? Moreover, how is it possible that we do not share these uncertainties with students, and unrealistically teach the chemical bond as if it were a closed topic? Is this not a kind of soft lying to the students, artificially shielding them from the reality of the chemical world and therefore leaving them unprepared?
A chemical bond course should begin by deconstructing it into its multiple definitions (Ayers et al., 2018; Hendry, 2008, 2022]; IUPAC, 2024a; Seifert, 2022b; Weisberg, 2008). We will describe here three approaches, their usefulness, their problems, and their unknowns. Considering their convenience, but also their limited range, they can also be labelled as chemical models (Schummer, 2014), as we shall discuss below. These are the energetic, structural, and orbital viewpoints of the bond.
3 The energetic approach to the chemical bond
IUPAC, the highest authority on anything that concerns chemical definitions, tells us that (IUPAC, 2024a):
“When forces acting between two atoms or groups of atoms lead to the formation of a stable independent molecular entity, a chemical bond is considered to exist between these atoms or groups.”
“The principal characteristic of a bond in a molecule is the existence of a region between the nuclei of constant potential contours that allows the potential energy to improve substantially by atomic contraction at the expense of only a small increase in kinetic energy.”
“Not only directed covalent bonds characteristic of organic compounds, but also bonds such as those existing between sodium cations and chloride anions in a crystal of sodium chloride or the bonds binding aluminium to six molecules of water in its environment, and even weak bonds that link two molecules of O2 into O4, are to be attributed to chemical bonds.”
Herein we will mostly digest part (a) of the IUPAC definition, characterizing the bond from the force.
3.1 A force acting between two atoms
Part (a) of IUPAC’s definition says that the chemical bond simply is a force that binds atoms. That is, if energy must be spent to stretch apart two atoms, then they are bonded. The definition does not expand on which kind of force, and therefore any force will be valid, from strong covalent to feeble London. Interestingly, the force must be exactly zero in a “stable independent molecular entity”, or the bonds would continuously shrink until reaching a singularity (see Figure 1). Mathematically, with V being the potential energy and r the distance between two atoms:
That is, a stable geometry involves a minimum in V, the potential energy surface, and a zero force. In fact, a molecule is defined as a multiatomic entity in a minimum of the potential energy surface (IUPAC, 2024b). Therefore, the definition of chemical bonds in a molecule in its ground state in terms of forces should be paradoxically expressed when the molecule is outside equilibrium, with stretched bonds. Note that the potential energy term here is not the Hamiltonian potential energy (obtained from the
![Figure 1:
Three examples of potential energies (solid lines, left axis in kJ mol−1) and forces acting in a bond [dashed lines, right axis in kJ mol−1 Å−1, see eq. (1) (Gaussian16, 2024)]. In a stationary geometry the forces are equal to zero by definition, since the potential energy of the molecule must be at its minimum. Note that by IUPAC’s definition the hydrogens of ethene would be bonded, as there is a force hindering their separation.](/document/doi/10.1515/cti-2024-0113/asset/graphic/j_cti-2024-0113_fig_001.jpg)
Three examples of potential energies (solid lines, left axis in kJ mol−1) and forces acting in a bond [dashed lines, right axis in kJ mol−1 Å−1, see eq. (1) (Gaussian16, 2024)]. In a stationary geometry the forces are equal to zero by definition, since the potential energy of the molecule must be at its minimum. Note that by IUPAC’s definition the hydrogens of ethene would be bonded, as there is a force hindering their separation.
It would be more accurate to define a chemical bond as a restoring force after they are stretched. The elongating and shrinking of the interatomic distance creates harmonic potentials (Figure 1), and therefore many people consider a chemical bond in a stable geometry by considering not the force (F in eq. (1)), but the force constant (f in eq. (2)) of the hypothetical spring between the atoms, i.e. the second derivative of the potential energy as a function of a small distortion of the bond distance (relaxing all other coordinates) from the molecule in its stable geometry (Brandhorst & Grunenberg, 2008; Kraka et al., 2020; Zhao et al., 2022).
One advantage of this definition is that it also gives us a measure of the bond strength. The larger f, the harder it is to stretch the bond, and therefore the stronger it will be. For instance, f for a single covalent C–C bond is around 400 N m−1, for the triple bond of N2 (one of the strongest bonds in nature) is ∼2,300 N m−1, and for non-covalent interactions it is usually lower than 100 N m−1.
The bigger drawback brought by this energy/force-based mathematical definition of the chemical bond is the fact that if you stretch any pair of atoms in any molecule you will observe a force that will attract them back to the optimal geometry. This is a necessary conclusion of defining the molecule as a minimum in the potential energy surface, which makes any geometrical distortion raise the molecular energy. It will happen even if these two atoms are far apart and removed by many ‘proper’ bonds. In other words, according to this definition, there is a pairwise attraction between all the atoms of a molecule, and therefore we will erroneously infer that all its atoms would be directly bonded, even the ones on opposite sides of the molecule. This would also be technically valid between the atoms of a large molecular cluster several nm apart, an idea that not many chemists would approve.
This is exemplified in Figure 1, which shows that if you elongate the distance between two hydrogens in ethene, the energy goes up and a force will be created that will push back the hydrogens to their lowest energy. However, it is inconceivable to think under our typical conceptualization that there is a proper chemical bond between these two hydrogens separated by two carbons. What would you call the ‘force’ between the hydrogens? Certainly not a covalent or ionic bond, not even a van der Waals interaction [although it might be considered a ‘through-bond interaction’ (Albright et al., 2013) or in certain cases a ‘mechanical bond’ (Bo, 2017)]. Without the carbons holding the hydrogens in their positions (as can be done in a model H2 dimer) they would repel each other, clearly indicating that there is no chemical bond between hydrogens in ethene, no matter what the stretching force tells us.
A possible solution to this problematic definition of the bond as a force might be to indicate that it works only when the atoms are already connected. However, this will fail due to a circular reasoning fallacy, as we will have to explain what makes this connection in the first place.
Another solution might be to model the pairwise interaction with a simplified system that removes all the molecular constraints; if they keep the attraction, then the bond was real. This can be done, for instance, by mimicking ethane with an H2 dimer (which would show that the H–H bond is spurious).
Another case is the short SiH–HSi interaction of the cage molecule shown in Figure 2 (Mandal et al., 2017; Mandal & Datta, 2020; Zong et al., 2013), a hydrogen-hydrogen ‘bond’ is kept in a very tight compartment in a kind of ‘iron maiden’ (Pascal, 2004) system. A strong force is required to separate both hydrogens, but that is an artificial effect brought by the constricted cage. If we release the constraints, which can be easily done with a silane dimer model (on the right of Figure 2), we will see that the bonding force is extrinsic to the hypothetical bond, completely beyond the H–H region. As a matter of fact, some orbital, energy decomposition, and topographical quantum mechanical methods indicate that there may be a weak force binding the hydrogens (Mandal et al., 2017), but it is completely overpowered by the Pauli repulsion without the external constraint. And so, what shall we say about this H–H interaction? Is there a chemical bond between the two hydrogens? Is a mathematical definition of the bond through the force enough to declare the existence of a bond? (Berson, 2008)

In this cage molecule (Mandal et al., 2017; Mandal & Datta, 2020; Zong et al., 2013), the confined internal hydrogens depict an ultra-short non-covalent interaction. Since it is in its optimal geometry, any intent to separate the hydrogens will require to employ a significant force, meaning that they have a chemical bond between them according to IUPAC’s definition. However, if we release the tension brought by the cage, such as with a silane dimer (but still keeping the symmetry), the bond is utterly broken (Gaussian16, 2024).
I will leave another strawman fallacy for the reader to analyse: Adenine and thymine are a base pair that are bound through two H-bonds (N–H and O–H), which are undisputed chemical bonds, even if they are not covalent. If we break the N–H bond by a rotation (lower part of Figure 3), the O–H bond is elongated to 2.02 Å. This implies that without the external force of the N–H bond, the O–H bond would be repulsive at the original bond length of 1.94 Å. If an attractive force is proof of a chemical bond, is a repulsive force proof of the lack of a chemical bond?

Adenine-thymine base pair. When breaking the N–H hydrogen bond the O–H bond is elongated, which indicates that there was a repulsive force between the oxygen and the hydrogen that was relaxed when the external forces were removed (Gaussian16, 2024). Does this mean that there was no chemical bond between the O and the H?
3.2 Everything goes
There are more issues with the energy/force definition of the chemical bond. One critical problem is that “This definition does not state what chemical bonds are; rather it states the conditions that hold when a chemical bond is considered to exist” (Seifert, 2022b). There is no solution to this problem when working only with raw energy values.
Parts (b) and (c) of IUPAC’s definition (see above) are also not free of challenges. The discussion about the potential and kinetic energies (part b of the definition) is a thorny and open topic of debate, as we try to comprehend which one of these two components of the bond energy is the real culprit of the chemical bonds.
Part (c) indicates that any type of interaction is valid to enlist it as a chemical bond, as long as it creates an attraction force: from the Na+ to Cl- ionic to the O2 dimer London interactions, everything goes. Wherever there is a force, no matter its nature or how weak it is, there is a bond. This is not only ambiguous, but also goes against our conventional understanding of the chemical bond. Do we have a chemical bond in every case of London forces? What about He2, a dimer 25,000 times weaker than the benzene dimer? What is the limit where we can safely say there is no bond anymore? (Berson, 2008)
The convenient energetic definition can represent and partly explain the bond, and it can predict where a bond should occur; but as seen above, its utility must be framed within limits. In this sense, we can consider it not only as a definition, but also as a model (Schummer, 2014) of the chemical bond. This has the advantage that students are used to working with chemical models of all types (VSEPR, ideal gases, crystal field, etc.), and therefore it is a point worthy of teaching. Another characteristic of models is that they can accommodate other models that may explain the same phenomenon with completely different mathematical or heuristic tools, each one of them having explanation and prediction power, as well as their own limits [or as the ironic aphorism says, “all models are wrong, but some are useful” (Box, 1979)]. Still, there are many models of the bond that do not work as definitions of the chemical bond (for example Lewis electron pair and Langmuir octet rule, hypervalency, σ-bonding/π-backbonding, etc.), and therefore we can conclude that most chemical definitions can work as models, but most models are not definitions. In this sense, with the understanding that they are also models, we will continue with two other definitions of the bond.
4 The structural approach
From a purely physical viewpoint, molecules are just a collection of nuclei and electrons in a geometry that minimizes their energy. However, we cannot be oblivious to the existence of a structural pattern (Seifert, 2022b) that shows something connecting the nuclei, mostly in a pairwise atom-to-atom fashion. This is implicit when drawing molecules with ball-and-stick or skeletal models, but it is also observed from computational and experimental images (see Figure 4). This is the structural [also called topological (Pendás & Contreras-García, 2023)] approach to the chemical bond. If we believe that electrons make the bond, the interatomic glue has to be the electron density. This idea is rooted in the Lewis electron pair, and was thoroughly developed through many models, possibly the most famous being the Quantum Theory of Atoms in Molecules [QTAIM (Bader, 1994; Matta & Boyd, 2007; Pendás & Contreras-García, 2023; Popelier, 2000)], which we will briefly discuss below.
![Figure 4:
Different pentacene depictions, from formular representations to computed or experimental images. All of them portray the atom-to-atom interaction as the default chemical bond pattern. Both the computed electronic and the experimental results substantiate the hypothesis that the chemical bond is a pairwise interaction based on a high electronic density. The AFM picture obtained in 2009 was revolutionary in terms of its technical achievement [adapted from (Gross et al., 2009), reprinted with permission from AAAS].](/document/doi/10.1515/cti-2024-0113/asset/graphic/j_cti-2024-0113_fig_004.jpg)
Different pentacene depictions, from formular representations to computed or experimental images. All of them portray the atom-to-atom interaction as the default chemical bond pattern. Both the computed electronic and the experimental results substantiate the hypothesis that the chemical bond is a pairwise interaction based on a high electronic density. The AFM picture obtained in 2009 was revolutionary in terms of its technical achievement [adapted from (Gross et al., 2009), reprinted with permission from AAAS].
Due to the atomic pairwise nature of the structural bonding, we can draw molecules as graphs (that is, points connected by lines) or define molecules just by simple text coding (like SMILES notation, where for example ‘C[C@H](C(=O)O)N’ is d-Alanine). The concept of pairwise bonding with some structural electronic glue is so intrinsic to the chemical bond paradigm that just by looking at a figure of connected lines we know exactly which molecule it is, and what are its basic properties.
The accumulation of electronic density as a bonding pattern, beyond its ubiquitousness, has a second advantage: it is an observable property [as in quantum mechanically observable (Ayers et al., 2018; Scherer et al., 2014)]. Opposed to the esoteric orbitals and wave functions, it can be experimentally checked, even if it is technically challenging [see Figure 4 (Gross et al., 2009)]. And contrary to the energy perspective of the bond, it is a phenomenon that can be observed inside the bond, directly between the involved atoms [although it was argued that bonding is not a submolecular phenomenon (Weisberg, 2008)].
At this point, it is important to understand how the electronic density is distributed in a molecule. First of all, most of the electron density will always be concentrated encircling the nuclei, since nothing lowers the potential energy more than having positive and negative particles close by (see the red small balls at the carbon positions in Figure 5). The structural chemical bond is, therefore, a matter mainly of the remaining valence electrons. As can be seen in Figure 5, due to the attraction to other nuclei and some kinetic energy interference, the valence electrons are concentrated between the atomic centers, creating the almost universal structural pattern of the chemical bond.

Phenanthrene electron density from 60 % to 99 % isodensity surfaces (the numbers indicate the amount of density enclosed in the volume). At 60 % (red spheres at the nuclei positions) it is possible to see that most of the electrons are in the carbon and hydrogen cores. At 70 % (orange surfaces) we start to see the atomic pairwise interactions. At 80 % and 90 % (yellow and blue) the structure of the bonds as a pairwise interatomic interaction is clear. At 99 % (purple) the pairwise bonding starts to be undefined. Superimposed are the QTAIM bonding paths (white lines) and the bond critical points (green dots). Between the upper central hydrogens there is an artefactual bond path and a critical point (AIMAll, 2019; Gaussian16, 2024).
We can draw a line of maximum electronic density between the atoms (white lines in Figure 5), and in the middle we can mark the point of minimal density in this coordinate (but maximum density perpendicularly, see green dots in Figure 5). According to QTAIM (Bader, 1994; Matta & Boyd, 2007; Pendás & Contreras-García, 2023; Popelier, 2000), these lines and points are the most faithful characteristic of the chemical bond. Moreover, cutting through these critical points we can distinguish one atom from the other in the molecule, providing a compelling (but controversial and hardly universally accepted) answer to the question of what an atom in a molecule is. Because of this, QTAIM is the pet model of the structural approach to the bond [there are many other techniques that also rely on mathematical acrobatics based on electronic densities such as ELF, NCI, IQA, LED, DORI, etc (Pendás & Contreras-García, 2023), each one of them with its own advantages].
As useful and appealing as it is, the structural approach is not flawless. Even after all its successes we can find cases of false positives (false bonds where there is no real bond) and false negatives (undetected bonds). Phenanthrene is one of the classical situations of the former. As shown in Figure 5, the density coming from the two upper central hydrogens creates a bonding path between them, giving the appearance of a weak but real bond according to the structural definition of the bond (see the purple surface, and the green critical point). The same effect would be seen if we squeeze two noble gas atoms, where a density build-up appears between them despite being a repulsive interaction. This is a simple mathematical effect of atomic densities brought together, as shown in Figure 6 with two s-type atomic functions. Even without any potential energy attraction or kinetic energy interference, there will be a low-density critical point, not much different from the H–H one in phenanthrene. This is problematic since the structural approach cannot be a fortuitous mathematical effect appearing when bringing atoms around.

Two close-by hydrogen-like radial functions. Even without any interference between the functions a critical point appears, giving the impression of a weak bond according to the structural definition.
Some of these difficulties can be eliminated by ad-hoc impositions to the definition of the structural bond, like asking not to contract the interatomic distances beyond their minimum energy if they were not constrained by the system. This condition is too restrictive, and resembles the problem of the atoms being forced into place that appeared in the energetic approach to the bond (see Figure 2).
A second possibility is to disregard the cases where the density between the atoms reaches negligible values. Filtering out the cases of too low density makes a lot of sense, but degrades the significance of this definition. Figure 7 shows the electronic density isosurfaces of NiCO4, where between the metal and the ligands the density is much smaller than the values between C–H and C–C bonds in phenanthrene (Figure 5). This corresponds to the fact that coordinative bonds tend to be weaker than traditional covalent ones. Even weaker, van der Waals interactions still have a QTAIM critical point that permits their detection, but their densities at those points is extremely low, almost negligible. This is problematic since the structural approach does not define the limit of a significant electronic build-up to consider it a bond.

Electron density at different isodensity surfaces for NiCO4, a classic complex. The electron density between the metal and the ligands already shows discontinuities at the 80 % surface (in yellow), a characteristic of a bond weaker than a covalent one (Gaussian16, 2024).
Electrostatic interactions also show critical bonds, but they do not show any meaningful superposition between the atomic electron densities spheres of interaction. For instance, NaF does not show any contact between the atomic densities even above the 99 % isosurface (Figure 8). This number is too high to justify the chemical bond as an electron density build-up model. In purely ionic bonds we must admit that whatever is making the bond, it is not something chemically tangible, but a physical coulombic force field. We may argue that opposed to covalent bonds in electrostatic interactions the ‘structure’ can be a build-up in the electric field (Weisberg, 2008), but that goes against the structural approach defined by electronic densities.

In the depicted NaF density isosurfaces not even at 99 % (purple surface) we see a chemical bond in the structural sense, despite being a strong interaction (Gaussian16, 2024).
In other words, the structural definition of the chemical bond is an excellent approach, but only for covalent interactions, a subset of the whole paradigm. These would be the limits of the model.
5 The orbital/valence/electron pair approach
5.1 The omnipresent but cryptic molecular orbital
Although the concept of a wave function is intrinsic to physics, the molecular orbital (MO) resides in the realm of chemistry (Fortin & Jaimes Arriaga, 2022; Weisberg, 2008), especially when building them from atomic orbitals (MO-LCAO). What is in the realm of philosophy is the question of whether the MO really exists or not, considering that it is not a quantum observable (Hettema, 2017; Krylov, 2020; Labarca & Lombardi, 2010; Mulder, 2011; Pham & Gordon, 2017; Scerri, 2000, 2001]; Truhlar et al., 2019; Zuo et al., 1999). Even if they may not be real, analyses such as photoelectron spectroscopy, Koopmans theorem, Fukui function, Dyson orbitals, and atomic resolution spectroscopy techniques indicate that molecules behave as if MOs were real. This is a fascinating debate, but here we will focus only on its utility as a quantitative and qualitative chemical tool (Jenkins, 2003).
The MOs approach to the chemical bond is simple. We count the orbitals, and if the number of bonding ones is larger than antibonding ones, then there is a chemical bond, as shown in the interaction diagram of Figure 9 for the triple bond of CO. Since each MO can hold up to two electrons, we have a direct connection with Lewis’ shared electron pair model. Can this work as a general definition of the chemical bond? Well, most chemistry books think so. It certainly has a lot of appeal in terms of explanatory and predictive power, much more than the previously discussed structural and energetic approaches, models that show the existence of bonds but have many difficulties explaining their raison d’être (Seifert, 2022b). For example, we can build the final orbitals from their fragments in interaction diagrams, allowing better comprehension of the interaction (Albright et al., 2013; Rauk, 2001). The structural and energetic definitions only consider what happens in the final product, or in other words, they work on the chemical bond, but not on the chemical bonding.

Interaction diagram for CO. There is a chemical bond since there are more bonding orbitals than antibonding ones. In this case, CO has a triple bond, with two covalent π and one dative (Nandi & Kozuch, 2020) σ bonding MOs.
MOs can be of very different styles, including canonical, Dyson, NTO, NBO, and others. These orbital schemes can be constructed from linear combinations of the other orbital styles, and they are all legitimate as long as they provide the correct electronic density (Krylov, 2020; Truhlar et al., 2019; Weinhold & Landis, 2005). Each one of these MO representations gives a useful depiction for a specific use (energetic, spectroscopic, reactivity, bonding, etc.), but they will disappoint if trying to use them for another property. For instance, despite what we learn in college, most canonical MOs do not provide any clear bonding pattern like the ‘clean’ CO orbitals of Figure 9. The canonical MOs (which are orthogonal to each other, and are the ones produced by most quantum chemistry methods and programs) are mostly delocalized all around the molecule, and therefore can be bonding between one pair of atoms, antibonding between another pair, and nonbonding for the rest (see Figure 10). While bond orders can be mathematically obtained from the canonical MOs (Mayer, 2007), these methods loose the simple visual connection between a bond and the orbitals. However, their localization can do the trick (see Figure 10), although the localization process may sound too artificial and begging the question for the taste of some physical chemistry purists, plus they lose any spectroscopical insight. In essence, each style of molecular orbital representation must be used for the purpose it was designed, for instance the orbitals localized in the bond can be used to define, quantify, and understand each bond (see the NBO and Boys localization in Figure 10).

Porphyrin molecular orbitals. The canonical MOs (HOMO and HOMO-1 in the figure) are completely delocalized, losing the chemical intuition that localization provides. The lower images show pairs of C–N bonding MOs in NBO and boys localization flavours. NBO produces the σ and π conventional MOs, boys generates the pauling-style ‘banana’ double bonds.
Here is another issue when taking the MO-based chemical bond in a simplistic way: According to the basic view of the model it is possible to formally have a single, double, triple, quadruple (Cotton & Harris, 1965), quintuple (Brynda et al., 2006; Frenking, 2005), or up to a sextuple (Frenking & Tonner, 2007; Roos et al., 2007) covalent bond; you might even have half a covalent bond (like in the 3-center-4 electron cases such as XeF2 or with F2 −), one and a half covalent bond (as in benzene), or another option if an odd number of electrons are involved. In this formal language, the covalent bond order is always a function of the number of electrons and the orbitals they occupy. However, this integer formality can disappear, producing fractional bond orders, for instance when there is superposition of localized orbitals, or when using multi-reference methods that result in fractional orbital occupancy. In other words, integer bond orders are only a formal idealization of real bonds. In contrast, when speaking about chemical bonds from their structural or energetic perspectives, we generally do not speak in terms of bond orders, but in terms of their quantitative strengths, just stating that a bond is weak or strong (Brandhorst & Grunenberg, 2008; Shaik et al., 2016).
But maybe the biggest disadvantage of the MO as a diagnostic of the chemical bond is, similar to the structural method, its difficulty in describing non-covalent bonds. MO theory as a computational approximation to the solution of Schrödinger equation can certainly compute the electrostatic and van der Waals bonding through ab initio or DFT methods, and there are orbital-based explanations for these interactions (Schneider et al., 2016; Truhlar, 2019). Nevertheless, no significant density between atoms implies that there are no significant bonding orbitals between them, limiting the model. We cannot say that orbitals make the bonds when there are no orbitals in these bonds.
5.2 Valence bond structures, the wave function in the image and likeness of the chemical bond
A final thought is briefly dedicated to valence bond theory (VB), a method with less computational use, less fame, and less flashy pictures compared to MO theory, but with a longer history (Pauling, 1960) and a better ability to define chemical bonds as chemists like to think about them (Shaik & Hiberty, 2008; Weisberg, 2008). Indeed, if à la Lewis we consider the shared electron pair as the definition of the chemical bond (Lewis, 1916), VB generates mathematical structures with a one-to-one correspondence to these pairs. If in MO theory the ground state MO of H2 is written as σ = sA + sB, in VB theory the covalent bond is written as ϕ HL = sA × sB (to write the physically correct expressions we must add spins and antisymmetrize the wave functions, but for the sake of simplicity we show here only its basic term). For MO we create an orbital to put the electrons as the positive linear combination of atomic orbitals (in this case the σ MO from the 2s AOs). In VB we create a covalent structure for that purpose, called ‘Heitler-London’ in honor of the authors of the first quantum description of a chemical bond (Heitler & London, 2000).
While with enough refinements VB and MO theory in the end converge to the same energy value, a great advantage of VB is that it can clearly define in conceptual and mathematical manner our most basal ideas of the bond in chemical terms, such as the bond order, resonance, covalent, and ionic characters. Therefore, if we qualitatively consider the chemical bond as shared electron pairs (or in Lewis’ words: “…to express this idea of chemical union in symbols I would suggest the use of a colon … to represent the two electrons which act as the connecting links between the two atoms” (Lewis, 1916)), valence bond theory provides its quantitative quantum mechanical justification. While VB computations are harder than MO theory computations, VB competes head-to-head in terms of explanatory and predictive power (Shaik & Hiberty, 2008). Still, even if high-level VB can compute weak interactions, it does not have any kind of van der Waals structure. In these cases, like with orbitals, we cannot say that shared electron pairs make the bonds when there are no pairs in these bonds. Nothing is perfect, but quoting Voltaire, “the best is the enemy of good”. And both VB and MO models are really good.
6 Conclusions
Ethics in chemical education is mostly focused on teaching the expected behavior of chemists in research and industry after graduation. However, there is an alternative, intrinsic facet of ethics in the approach in which lecturers teach chemistry, specifically in the honesty they impart regarding the description of the knowledge of the field. This is a responsibility seldom faced in class, where uncertainties are disregarded, and knowledge is imparted as closed topics (with the exception of quantum mechanics, where ignorance and uncertainties are embedded in the field). Under this idea, herein we described the biggest paradigm in chemistry, the chemical bond. We briefly reviewed three possible ways to define it:
On the basis of energies, checking how hard it is to separate atoms.
On the basis of the electron density, which shows a structural bond between atoms.
On the basis of orbitals or VB structures, depicting shared electron pair type of wave functions between atoms.
The three definitions are useful and insightful, but also lacking and different between each other (Ayers et al., 2018; Hendry, 2008, 2022]; Seifert, 2022b; Weisberg, 2008). Many, if not most supposedly clear chemical concepts are actually ‘fuzzy’ or ill-defined (Frenking & Krapp, 2006; Gonthier et al., 2012; Grunenberg, 2017), and notoriously difficult to be demarcated in a foolproof definition; and the first hard-to-state definition is the chemical bond itself. Moreover, in all definitions we can find false positive and false negative cases if we look deep enough. As discussed above, maybe instead of calling them definitions we should call them models, since they are not right or wrong, but useful between their limits (Schummer, 2014). The chemical bond is a complex, multidimensional entity with many unknowns. Defining it in one sentence cannot possibly do justice to it, but we can still have simple, partial, working definitions that can guide us to see the pattern (Seifert, 2022b). The chemical bond is real, but since it also is a complex, multifaceted concept, any single definition will necessarily be a restricting interpretation. This dichotomy should be discussed with students as part of the puzzles in the chemical world and in the scientific method.
Or maybe we can just disregard any definition and stick to the IKIWISI principle, rooted in the famous phrase of US Supreme Court Justice Potter Stewart, accepting that for the chemical bond… “I shall not attempt to define the kinds of material I understand to be embraced within that shorthand description, and perhaps I could never succeed in intelligibly doing so. But I know it when I see it.” (Wikipedia, 2023).
Funding source: PAZY Foundation
Award Identifier / Grant number: ID(416)-2023
-
Research ethics: Not applicable.
-
Informed consent: Not applicable.
-
Author contributions: The author has accepted responsibility for the entire content of this manuscript and approved its submission.
-
Use of Large Language Models, AI and Machine Learning Tools: None declared.
-
Conflict of interest: The author states no conflict of interest.
-
Research funding: This research is supported by the PAZY foundation [ID(416) – 2023].
-
Data availability: Not applicable.
References
AIMAll (2019). AIMAll (Version 19.10.12). Overland Park KS, USA: TK Gristmill Software. 2019 (aim.tkgristmill.com).Suche in Google Scholar
Albright, T. A., Burdett, J. K., & Whangbo, M.-H. (2013). Orbital interactions in chemistry (2nd ed.). Wiley.10.1002/9781118558409Suche in Google Scholar
Ayers, P. W., Fias, S., & Heidar-Zadeh, F. (2018). The axiomatic approach to chemical concepts. Computational and Theoretical Chemistry, 1142, 83–87. https://doi.org/10.1016/j.comptc.2018.09.006.Suche in Google Scholar
Bader, R. F. (1994). Atoms in molecules: A quantum theory. Clarendon Press.Suche in Google Scholar
Berson, J. A. (2008). Molecules with very weak bonds: The edge of covalency. Philosophy of Science, 75(5), 947–957. https://doi.org/10.1086/594537.Suche in Google Scholar
Bo, G. D. (2017). Mechanochemistry of the mechanical bond. Chemical Science, 9(1), 15–21. https://doi.org/10.1039/C7SC04200K.Suche in Google Scholar PubMed PubMed Central
Box, G. E. P. (1979). Robustness in the strategy of scientific model building. In R. L. Launer & G. N. Wilkinson (Eds.), Robustness in Statistics (pp. 201–236). Academic Press.10.1016/B978-0-12-438150-6.50018-2Suche in Google Scholar
Brandhorst, K., & Grunenberg, J. (2008). How strong is it? The interpretation of force and compliance constants as bond strength descriptors. Chemical Society Reviews, 37(8), 1558–1567. https://doi.org/10.1039/B717781J.Suche in Google Scholar PubMed
Brynda, M., Gagliardi, L., Widmark, P.-O., Power, P. P., & Roos, B. O. (2006). A quantum chemical study of the quintuple bond between two chromium centers in [PhCrCrPh]: Trans-bent versus linear geometry. Angewandte Chemie International Edition, 45(23), 3804–3807. https://doi.org/10.1002/anie.200600110.Suche in Google Scholar PubMed
Burrows, N. L., & Mooring, S. R. (2015). Using concept mapping to uncover students’ knowledge structures of chemical bonding concepts. Chemistry Education: Research and Practice, 16(1), 53–66. https://doi.org/10.1039/C4RP00180J.Suche in Google Scholar
Çalik, M., Ültay, N., Bağ, H., & Ayas, A. (2024). A meta-analysis of effectiveness of chemical bonding-based intervention studies in improving academic performance. Chemistry Education: Research and Practice, 25, 506–523. https://doi.org/10.1039/D3RP00258F.Suche in Google Scholar
Clark, T., Politzer, P., & Murray, J. S. (2022). Interpreting the variations in the kinetic and potential energies in the formation of a covalent bond. Physical Chemistry Chemical Physics, 24, 12116–12120. https://doi.org/10.1039/D2CP01529C.Suche in Google Scholar
Coll, R. K., & Taylor, N. (2001). Alternative conceptions of chemical bonding held by upper secondary and tertiary students. Research in Science & Technological Education, 19(2), 171–191. https://doi.org/10.1080/02635140120057713.Suche in Google Scholar
Cotton, F. A., & Harris, C. B. (1965). The crystal and molecular structure of dipotassium octachlorodirhenate(III) dihydrate, K2[Re2Cl8]2H2O. Inorganic Chemistry, 4(3), 330–333. https://doi.org/10.1021/ic50025a015.Suche in Google Scholar
Coulson, C. A. (1955). The contributions of wave mechanics to chemistry. Journal of the Chemical Society (Resumed), 2069–2084. https://doi.org/10.1039/JR9550002069.Suche in Google Scholar
Crabtree, H. R. (2017). Hypervalency, secondary bonding and hydrogen bonding: Siblings under the skin. Chemical Society Reviews, 46, 1720–1729. https://doi.org/10.1039/C6CS00688D.Suche in Google Scholar PubMed
Croft, M., & de Berg, K. (2014). From common sense concepts to scientifically conditioned concepts of chemical bonding: An historical and textbook approach designed to address learning and teaching issues at the secondary school level. Science & Education, 23(9), 1733–1761. https://doi.org/10.1007/s11191-014-9683-0.Suche in Google Scholar
Deveson, A., Cremer, D., Frenking, G., Piris, M., & Shaik, S. (2016). Why does C2 cause so many problems? ChemViews. https://doi.org/10.1002/chemv.201600022.Suche in Google Scholar
Dirac, P. A. M. (1929). Quantum mechanics of many-electron systems. In Proceedings of the Royal Society of London – Series A: Containing Papers of a Mathematical and Physical Character. Vol. 123(792) (pp. 714–733).10.1098/rspa.1929.0094Suche in Google Scholar
Dulmen, T. H. H., van Visser, T. C., Coenders, F. G. M., Pepin, B., & McKenney, S. (2023). Learning to teach chemical bonding: A framework for preservice teacher educators. Chemistry Education: Research and Practice, 24(3), 896–913. https://doi.org/10.1039/D2RP00049K.Suche in Google Scholar
Erduran, S., & Kaya, E. (2019). Philosophy of chemistry and chemistry education. In S. Erduran & E. Kaya (Eds.), Transforming Teacher Education Through the Epistemic Core of Chemistry: Empirical Evidence and Practical Strategies (pp. 1–24). Springer International Publishing.10.1007/978-3-030-15326-7_1Suche in Google Scholar
Fortin, S., & Jaimes Arriaga, J. A. (2022). About the nature of the wave function and its dimensionality: The case of quantum chemistry. In O. Lombardi, J. C. Martínez González & S. Fortin (Eds.), Philosophical Perspectives in Quantum Chemistry (pp. 203–216). Springer International Publishing.10.1007/978-3-030-98373-4_9Suche in Google Scholar
Frank, H., Campanella, L., Dondi, F., Mehlich, J., Leitner, E., & Rossi, G. (2011). Ethics, chemistry, and education for sustainability. Angewandte Chemie International Edition, 50(37), 8482–8490. https://doi.org/10.1002/anie.201007599.Suche in Google Scholar PubMed
Frenking, G. (2005). Building a quintuple bond. Science, 310(5749), 796–797. https://doi.org/10.1126/science.1120281.Suche in Google Scholar PubMed
Frenking, G. (2023). Heretical thoughts about the present understanding and description of the chemical bond. Molecular Physics, 121(9–10), e2110168. https://doi.org/10.1080/00268976.2022.2110168.Suche in Google Scholar
Frenking, G., & Krapp, A. (2006). Unicorns in the world of chemical bonding models. Journal of Computational Chemistry, 28(1), 15–24. https://doi.org/10.1002/jcc.20543.Suche in Google Scholar PubMed
Frenking, G., & Tonner, R. (2007). The six-bond bound. Nature, 446(7133), 276–277. https://doi.org/10.1038/446276a.Suche in Google Scholar PubMed
Frenking, G. & Shaik, S. (Eds.). (2014). The Chemical Bond: Fundamental Aspects of Chemical Bonding (1st ed.). Wiley.10.1002/9783527664658Suche in Google Scholar
Garritz, A. (2013). Teaching the philosophical interpretations of quantum mechanics and quantum chemistry through controversies. Science & Education, 22(7), 1787–1807. https://doi.org/10.1007/s11191-012-9444-x.Suche in Google Scholar
Gaussian16. (2024). Carried out with M06-2x/6-311G(d) with Gaussian16.Suche in Google Scholar
Gonthier, J. F., Steinmann, S. N., Wodrich, M. D., & Corminboeuf, C. (2012). Quantification of “fuzzy” chemical concepts: A computational perspective. Chemical Society Reviews, 41(13), 4671. https://doi.org/10.1039/c2cs35037h.Suche in Google Scholar PubMed
Gross, L., Mohn, F., Moll, N., Liljeroth, P., & Meyer, G. (2009). The chemical structure of a molecule resolved by atomic force microscopy. Science, 325(5944), 1110–1114. https://doi.org/10.1126/science.1176210.Suche in Google Scholar PubMed
Grossman, M. I. (2017). John Dalton and the origin of the atomic theory: Reassessing the influence of bryan higgins. The British Journal for the History of Science, 50(4), 657–676. https://doi.org/10.1017/S0007087417000851.Suche in Google Scholar PubMed
Grunenberg, J. (2017). Ill-defined chemical concepts: The problem of quantification. International Journal of Quantum Chemistry, 117(9), e25359. https://doi.org/10.1002/qua.25359.Suche in Google Scholar
Heitler, W., & London, F. (2000). Interaction between neutral atoms and homopolar binding according to quantum mechanics. In H. Hettema (Ed.), World Scientific Series in 20th Century Chemistry. Vol. 8, (pp. 140–155). World Scientific.10.1142/9789812795762_0009Suche in Google Scholar
Hendry, R. F. (2008). Two conceptions of the chemical bond. Philosophy of Science, 75(5), 909–920. https://doi.org/10.1086/594534.Suche in Google Scholar
Hendry, R. F. (2022). Quantum mechanics and molecular structure. In O. Lombardi, J. C. Martínez González & S. Fortin (Eds.), Philosophical Perspectives in Quantum Chemistry (pp. 147–172). Springer International Publishing.10.1007/978-3-030-98373-4_7Suche in Google Scholar
Hettema, H. (2017). Atoms, chemical atoms and orbitals as chemical objects. In H. Hettema (Ed.), The Union of Chemistry and Physics: Linkages, Reduction, Theory Nets and Ontology (pp. 261–271). Springer International Publishing.10.1007/978-3-319-60910-2_10Suche in Google Scholar
Hilton, A., & Nichols, K. (2011). Representational classroom practices that contribute to students’ conceptual and representational understanding of chemical bonding. International Journal of Science Education, 33(16), 2215–2246. https://doi.org/10.1080/09500693.2010.543438.Suche in Google Scholar
Hunter, K. H., Rodriguez, J.-M. G., & Becker, N. M. (2022). A review of research on the teaching and learning of chemical bonding. Journal of Chemical Education, 99(7), 2451–2464. https://doi.org/10.1021/acs.jchemed.2c00034.Suche in Google Scholar
IUPAC (2024a). Iupac – chemical bond (CT07009). https://doi.org/10.1351/goldbook.CT07009.Suche in Google Scholar
IUPAC (2024b). Iupac – molecule (M04002). https://doi.org/10.1351/goldbook.M04002.Suche in Google Scholar
Jenkins, Z. (2003). Do you need to believe in orbitals to use them? Realism and the autonomy of chemistry. Philosophy of Science, 70(5), 1052–1062. https://doi.org/10.1086/377388.Suche in Google Scholar
Kaldaras, L., Akaeze, H. O., & Krajcik, J. (2024). Developing and validating an Next Generation Science Standards-aligned construct map for chemical bonding from the energy and force perspective. Journal of Research in Science Teaching, 61(7), 1689–1726. https://doi.org/10.1002/tea.21906.Suche in Google Scholar
Keinan, E. (2022). The chemist’s oath. Chemistry International, 44(4), 2–3. https://doi.org/10.1515/ci-2022-0401.Suche in Google Scholar
Kovac, J. (1996). Scientific ethics in chemical education. Journal of Chemical Education, 73(10), 926. https://doi.org/10.1021/ed073p926.Suche in Google Scholar
Kovac, J. (2006). Professional ethics in science. In D. Baird, E. Scerri & L. McIntyre (Eds.), Philosophy Of Chemistry: Synthesis of a New Discipline (pp. 157–169). Netherlands: Springer.Suche in Google Scholar
Kozuch, S. (2022). How scientific is chemistry? Israel Journal of Chemistry, 62, e202000112. https://doi.org/10.1002/ijch.202000112.Suche in Google Scholar
Kraka, E., Zou, W., & Tao, Y. (2020). Decoding chemical information from vibrational spectroscopy data: Local vibrational mode theory. WIREs Computational Molecular Science, 10(5), e1480. https://doi.org/10.1002/wcms.1480.Suche in Google Scholar
Krylov, A. I. (2020). From orbitals to observables and back. The Journal of Chemical Physics, 153(8), 080901. https://doi.org/10.1063/5.0018597.Suche in Google Scholar PubMed
Labarca, M., & Lombardi, O. (2010). Why orbitals do not exist? Foundations of Chemistry, 12(2), 149–157. https://doi.org/10.1007/s10698-010-9086-5.Suche in Google Scholar
Lavi, R., Shwartz, G., & Dori, Y. J. (2019). Metacognition in chemistry education: A literature review. Israel Journal of Chemistry, 59(6–7), 583–597. https://doi.org/10.1002/ijch.201800087.Suche in Google Scholar
Lee, R., & Cheng, M. M. W. (2014). The relationship between teaching and learning of chemical bonding and structures. In C. Bruguière, A. Tiberghien & P. Clément (Eds.), Topics and Trends in Current Science Education: 9th ESERA Conference Selected Contributions, Netherlands. (pp. 403–417). Springer.10.1007/978-94-007-7281-6_25Suche in Google Scholar
Legodi, M. A. (2021). Are our students learning and understanding chemistry as intended? Investigating the level of prior knowledge of UNIVEN students for the second year inorganic chemistry module. In L. Mammino & J. Apotheker (Eds.), Research in Chemistry Education (pp. 69–83). Springer International Publishing.10.1007/978-3-030-59882-2_5Suche in Google Scholar
Levine, D. S., & Head-Gordon, M. (2020). Clarifying the quantum mechanical origin of the covalent chemical bond. Nature Communications, 11, 4893. https://doi.org/10.1038/s41467-020-18670-8.Suche in Google Scholar PubMed PubMed Central
Levy Nahum, T., Mamlok‐Naaman, R., Hofstein, A., & Taber, K. S. (2010). Teaching and learning the concept of chemical bonding. Studies in Science Education, 46(2), 179–207. https://doi.org/10.1080/03057267.2010.504548.Suche in Google Scholar
Lewis, G. N. (1916). The atom and the molecule. Journal of the American Chemical Society, 38(4), 762–785. https://doi.org/10.1021/ja02261a002.Suche in Google Scholar
Mandal, N., & Datta, A. (2020). Molecular designs for expanding the limits of ultralong C–C bonds and ultrashort H-H non-bonded contacts. Chemical Communications, 56(98), 15377–15386. https://doi.org/10.1039/D0CC06690G.Suche in Google Scholar
Mandal, N., Pratik, S. M., & Datta, A. (2017). Exploring ultrashort hydrogen–hydrogen nonbonded contacts in constrained molecular cavities. The Journal of Physical Chemistry B, 121(4), 825–834. https://doi.org/10.1021/acs.jpcb.6b12391.Suche in Google Scholar PubMed
Mann, M. K. (2017). The right place and the right time: Incorporating ethics into the undergraduate biochemistry curriculum. In Liberal Arts Strategies for the Chemistry Classroom. Vol. 1266. (pp. 45–70). American Chemical Society.10.1021/bk-2017-1266.ch004Suche in Google Scholar
Matta, C. F. & Boyd, R. J. (Eds.). (2007). The quantum theory of atoms in molecules: From solid state to DNA and drug design. Wiley VCH.10.1002/9783527610709Suche in Google Scholar
Mayer, I. (2007). Bond order and valence indices: A personal account. Journal of Computational Chemistry, 28(1), 204–221. https://doi.org/10.1002/jcc.20494.Suche in Google Scholar PubMed
McClure, C. P., & Lucius, A. L. (2010). Implementing and evaluating a chemistry course in chemical ethics and civic responsibility. Journal of Chemical Education, 87(11), 1171–1175. https://doi.org/10.1021/ed1005135.Suche in Google Scholar
McGuffin, V. L. (2008). Teaching research ethics: It takes more than good science to make a good scientist. Analytical and Bioanalytical Chemistry, 390(5), 1209–1215. https://doi.org/10.1007/s00216-007-1800-3.Suche in Google Scholar PubMed
Mehlich, J., Moser, F., Van Tiggelen, B., Campanella, L., & Hopf, H. (2017). The ethical and social dimensions of chemistry: Reflections, considerations, and clarifications. Chemistry – A European Journal, 23(6), 1210–1218. https://doi.org/10.1002/chem.201605259.Suche in Google Scholar PubMed
Mulder, P. (2011). Are orbitals observable? HYLE – An International Journal for the Philosophy of Chemistry, 17(1), 24–35.Suche in Google Scholar
Nahum, T. L., Mamlok-Naaman, R., & Hofstein, A. (2013). Teaching and learning of the chemical bonding concept: Problems and some pedagogical issues and recommendations. In G. Tsaparlis & H. Sevian (Eds.), Concepts of Matter in Science Education (pp. 373–390). Netherlands: Springer.10.1007/978-94-007-5914-5_18Suche in Google Scholar
Nahum, T. L., Mamlok-Naaman, R., Hofstein, A., & Krajcik, J. (2007). Developing a new teaching approach for the chemical bonding concept aligned with current scientific and pedagogical knowledge. Science Education, 91(4), 579–603. https://doi.org/10.1002/sce.20201.Suche in Google Scholar
Nandi, A., & Kozuch, S. (2020). History and future of dative bonds. Chemistry – A European Journal, 26, 759–772. https://doi.org/10.1002/chem.201903736.Suche in Google Scholar PubMed
Neurath, O. (1973). Anti-spengler. In O. Neurath, M. Neurath & R. S. Cohen (Eds.), Empiricism and Sociology (pp. 158–213). Netherlands: Springer.10.1007/978-94-010-2525-6_6Suche in Google Scholar
Nordholm, S., & Bacskay, G. B. (2020). The basics of covalent bonding in terms of energy and dynamics. Molecules, 25(11). Article 11 https://doi.org/10.3390/molecules25112667.Suche in Google Scholar PubMed PubMed Central
Pascal, R. A. (2004). Molecular “iron maidens”: Ultrashort nonbonded contacts in cyclophanes and other crowded molecules. European Journal of Organic Chemistry, 2004(18), 3763–3771. https://doi.org/10.1002/ejoc.200400183.Suche in Google Scholar
Pauling, L. (1960). The nature of the chemical bond (3 ed., 17. print). Cornell Univ. Press.Suche in Google Scholar
Pendás, Á. M., & Contreras-García, J. (2023). Topological approaches to the chemical bond. Springer.Suche in Google Scholar
Pham, B. Q., & Gordon, M. S. (2017). Can orbitals really Be observed in scanning tunneling microscopy experiments? The Journal of Physical Chemistry A, 121(26), 4851–4852. https://doi.org/10.1021/acs.jpca.7b05789.Suche in Google Scholar PubMed
Pimple, K. D. (2002). Six domains of research ethics. Science and Engineering Ethics, 8(2), 191–205. https://doi.org/10.1007/s11948-002-0018-1.Suche in Google Scholar PubMed
Popelier, P. (2000). Atoms in molecules: An introduction. Prentice Hall.Suche in Google Scholar
Rauk, A. (2001). Orbital interaction theory of organic chemistry (2nd ed.). Wiley-Interscience.10.1002/0471220418Suche in Google Scholar
Roos, B. O., Borin, A. C., & Gagliardi, L. (2007). Reaching the maximum multiplicity of the covalent chemical bond. Angewandte Chemie International Edition, 46(9), 1469–1472. https://doi.org/10.1002/anie.200603600.Suche in Google Scholar PubMed
Ryan, H. (2024). Collective bonding continues to divide opinion. In Chemistry World. https://www.chemistryworld.com/news/collective-bonding-continues-to-divide-opinion/4018967.article.Suche in Google Scholar
Scerri, E. R. (2000). Have orbitals really been observed? Journal of Chemical Education, 77(11), 1492. https://doi.org/10.1021/ed077p1492.Suche in Google Scholar
Scerri, E. R. (2001). The recently claimed observation of atomic orbitals and some related philosophical issues. Philosophy of Science, 68(S3), S76–S88. https://doi.org/10.1086/392899.Suche in Google Scholar
Schelble, S. M. & Elkins, K. M. (Eds.). (2021), International ethics in chemistry: Developing common values across cultures. American Chemical Society.10.1021/bk-2021-1401Suche in Google Scholar
Scherer, W., Fischer, A., & Eickerling, G. (2014). The experimental density perspective of chemical bonding. In The Chemical Bond (pp. 309–344). John Wiley & Sons, Ltd.10.1002/9783527664696.ch9Suche in Google Scholar
Schneider, W. B., Bistoni, G., Sparta, M., Saitow, M., Riplinger, C., Auer, A. A. & Neese, F. (2016). Decomposition of intermolecular interaction energies within the local pair natural orbital coupled cluster framework. Journal of Chemical Theory and Computation, 12(10), 4778–4792. https://doi.org/10.1021/acs.jctc.6b00523.Suche in Google Scholar PubMed
Schummer, J. & Børsen, T. (Eds.). (2022), Ethics of chemistry: From poison gas to climate engineering (Paperback edition). World Scientific.Suche in Google Scholar
Schummer, J. (2014). The preference of models over laws of nature in chemistry. European Review, 22(S1), S87–S101. https://doi.org/10.1017/S1062798713000781.Suche in Google Scholar
Seifert, V. A. (2022a). In search of the chemical bond. In Chemistry world. https://www.chemistryworld.com/opinion/in-search-of-the-chemical-bond/4015825.article.Suche in Google Scholar
Seifert, V. A. (2022b). The chemical bond is a real pattern. Philosophy of Science, 1–47. https://doi.org/10.1017/psa.2022.17.Suche in Google Scholar
Shaik, S., Danovich, D., Braida, B., & Hiberty, P. C. (2016). The quadruple bonding in C2 reproduces the properties of the molecule. Chemistry – A European Journal, 22(12), 4116–4128. https://doi.org/10.1002/chem.201600011.Suche in Google Scholar PubMed
Shaik, S., & Hiberty, P. C. (2008). A chemist’s guide to valence bond theory. Wiley-Interscience.10.1002/9780470192597Suche in Google Scholar
Solà, M. (2024). Reflexiones en torno a la gestión de los grupos de investigación. Anales de Química de La RSEQ, 120(3), 131. https://doi.org/10.62534/rseq.aq.1995.Suche in Google Scholar
Sterling, A. J., Levine, D. S., Aldossary, A., & Head-Gordon, M. (2024). Chemical bonding and the role of node-induced electron confinement. Journal of the American Chemical Society, 146(14), 9532–9543. https://doi.org/10.1021/jacs.3c10633.Suche in Google Scholar PubMed
Stewart, I. & Cohen, J. (Eds.). (1997), Figments of reality: The evolution of the curious mind. Cambridge University Press.10.1017/CBO9780511541384Suche in Google Scholar
Stojanovska, M. (2024). Integrating ethics and democratic principles in chemistry education: A case study. Chemistry Teacher International, 6(4), 373–383. https://doi.org/10.1515/cti-2024-0010.Suche in Google Scholar
Sweeting, L. M. (1999). Ethics in science for undergraduate students. Journal of Chemical Education, 76(3), 369. https://doi.org/10.1021/ed076p369.Suche in Google Scholar
Tarannam, N., Shukla, R., & Kozuch, S. (2021). Yet another perspective on hole interactions. Physical Chemistry Chemical Physics, 23(36), 19948–19963. https://doi.org/10.1039/D1CP03533A.Suche in Google Scholar PubMed
Truhlar, D. G. (2019). Dispersion forces: Neither fluctuating nor dispersing. Journal of Chemical Education, 96(8), 1671–1675. https://doi.org/10.1021/acs.jchemed.8b01044.Suche in Google Scholar
Truhlar, D. G., Hiberty, P. C., Shaik, S., Gordon, M. S., & Danovich, D. (2019). Orbitals and the Interpretation of Photoelectron Spectroscopy and (e,2e) Ionization Experiments. Angewandte Chemie International Edition, 58(36), 12332–12338. https://doi.org/10.1002/anie.201904609.Suche in Google Scholar PubMed
Tsaparlis, G. (1997). Atomic and molecular structure in chemical education: A critical analysis from various perspectives of science education. Journal of Chemical Education, 74(8), 922. https://doi.org/10.1021/ed074p922.Suche in Google Scholar
Tsaparlis, G., Pappa, E. T., & Byers, B. (2018). Teaching and learning chemical bonding: Research-based evidence for misconceptions and conceptual difficulties experienced by students in upper secondary schools and the effect of an enriched text. Chemistry Education: Research and Practice, 19(4), 1253–1269. https://doi.org/10.1039/C8RP00035B.Suche in Google Scholar
Ünal, S., Çalık, M., Ayas, A., & Coll, R. K. (2006). A review of chemical bonding studies: Needs, aims, methods of exploring students’ conceptions, general knowledge claims and students’ alternative conceptions. Research in Science & Technological Education, 24(2), 141–172. https://doi.org/10.1080/02635140600811536.Suche in Google Scholar
Venkataraman, B. (2017). Emphasizing the significance of electrostatic interactions in chemical bonding. Journal of Chemical Education, 94(3), 296–303. https://doi.org/10.1021/acs.jchemed.6b00409.Suche in Google Scholar
Weinhold, F., & Landis, C. R. (2005). Valency and bonding: A natural bond orbital donor-acceptor perspective. Cambridge University Press.Suche in Google Scholar
Weisberg, M. (2008). Challenges to the structural conception of chemical bonding. Philosophy of Science, 75(5), 932–946. https://doi.org/10.1086/594536.Suche in Google Scholar
Wikipedia (2023). I know it when i see it. In Wikipedia. https://en.wikipedia.org/wiki/I_know_it_when_I_see_it.10.2139/ssrn.4524529Suche in Google Scholar
Yayon, M., Mamlok-Naaman, R., & Fortus, D. (2012). Characterizing and representing student’s conceptual knowledge of chemical bonding. Chemistry Education: Research and Practice, 13(3), 248–267. https://doi.org/10.1039/C0RP90019B.Suche in Google Scholar
Zhao, L., Zhi, M., & Frenking, G. (2022). The strength of a chemical bond. International Journal of Quantum Chemistry, 122(8), e26773. https://doi.org/10.1002/qua.26773.Suche in Google Scholar
Zohar, A. R., & Levy, S. T. (2019). Students’ reasoning about chemical bonding: The lacuna of repulsion. Journal of Research in Science Teaching, 56(7), 881–904. https://doi.org/10.1002/tea.21532.Suche in Google Scholar
Zong, J., Mague, J. T., & Pascal, R. A.Jr. (2013). Exceptional steric congestion in an in,in-Bis(hydrosilane). Journal of the American Chemical Society, 135(36), 13235–13237. https://doi.org/10.1021/ja407398w.Suche in Google Scholar PubMed
Zuo, J. M., Kim, M., O’Keeffe, M., & Spence, J. C. H. (1999). Direct observation of d-orbital holes and Cu–Cu bonding in Cu2O. Nature, 401(6748), 49–52. https://doi.org/10.1038/43403.Suche in Google Scholar
© 2024 the author(s), published by De Gruyter, Berlin/Boston
This work is licensed under the Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International License.
Artikel in diesem Heft
- Frontmatter
- Editorial
- The teaching of ethics and core values in chemistry education
- Special Issue Papers
- Teaching responsible chemistry: a challenge-based learning framework for the implementation of RRI courses in tertiary chemistry education
- A teaching module in research integrity and ethics for university students based on the IUPAC living-code approach
- Implementation of the course “good chemistry: methodological, ethical and social implications” – a case study
- Integrating ethics and democratic principles in chemistry education: a case study
- AI for chemistry teaching: responsible AI and ethical considerations
- From forensic chemistry: an educational experience
- Interactive ethics teaching for students of chemistry
- Ethics within chemistry education: options, challenges and perspectives
- Chemical ethics practices in HEBUST of China
- Do we know the chemical bond? A case for the ethical teaching of undefined paradigms
Artikel in diesem Heft
- Frontmatter
- Editorial
- The teaching of ethics and core values in chemistry education
- Special Issue Papers
- Teaching responsible chemistry: a challenge-based learning framework for the implementation of RRI courses in tertiary chemistry education
- A teaching module in research integrity and ethics for university students based on the IUPAC living-code approach
- Implementation of the course “good chemistry: methodological, ethical and social implications” – a case study
- Integrating ethics and democratic principles in chemistry education: a case study
- AI for chemistry teaching: responsible AI and ethical considerations
- From forensic chemistry: an educational experience
- Interactive ethics teaching for students of chemistry
- Ethics within chemistry education: options, challenges and perspectives
- Chemical ethics practices in HEBUST of China
- Do we know the chemical bond? A case for the ethical teaching of undefined paradigms