Teenage lust and star-crossed electrons

A new study reports a creative approach to modelling the atom motivated by a love story


Keith S. Taber


Perhaps it would be better not to introduce an orbital model until we feel learners are ready to appreciate the quantum jump from concentric orbits to fuzzy, overlapping, infinitely-extended patterns of electronic probability, and the associated complex patterns of energy levels they generate.


A scene from the play 'Romeo and Juliet'
"Grade: B-.
Comment: Your model of the heteronuclear molecule of Romeo-Juliet was creative and aesthetically pleasing, but it was inconsistent because you used rope to stand for the covalent bond when you are representing electrons with apples." (Image by Николай Оберемченко from Pixabay)


The science curriculum contains a good deal of abstract material that is both challenging, and – sadly – not always found intrinsically interesting, to many learners. The teacher has to find what can 'make the unfamiliar familiar', something I have written quite a lot about on this site.

Read about teaching as making the unfamiliar familiar

Modelling 'the' atom

One such abstract topic is the structure of 'the' atom 1 – an area where learners will likely come across multiple models and diverse representations, and where what is being modelled and represented (as a quanticle – a quantum object) simply cannot be adequately represented concretely. Given that, it is hardly surprising that often even keen and capable learners show alternative conceptions in this topic (Taber, 2002 [Download paper]).

I was therefore intrigued by a recent research paper that described an approach to progressing learners' ideas about atomic structure by asking them to engage with a story. Narrative is a recognised way of helping make the unfamiliar familiar, and here a story was referenced that is familiar to many people: that of Shakespeare's 'star-crossed lovers': Romeo and Juliet.

So, in the storyline, electrons were named after characters from the tragic tale. It is common to relate abstract chemical ideas to social relations (chemistry uses such metaphors as 'sharing electrons', 'nucleus loving' species, reagent species that 'attack' other molecules, and substances that 'compete') – but this does risk the anthropomorphism (that is, treating non-human entities as if they have human qualities) actually confusing learners.

Read about anthropomorphism and science

That is, molecules and ions, and nuclei and electrons are not like people, and do not think or have desires, and so they do not act from motivations such as love or hate or jealousy…

Perhaps this seems SO OBVIOUS that only the weakest student could possibly get confused and think otherwise?

But I know from my own research (e.g., Taber & Watts, 1996 [download paper]) that actually even studious, intelligent learners can come to habitually use anthropomorphic language without noticing that they are explaining chemistry in terms that would only make sense if atoms and molecules and ions and electrons did have preferences, and could think for themselves, and did act accordingly!

Atoms can not care about anything – so they do not care about how many electrons they have, and they never deliberately do anything in order to obtain full shells or octets (as they cannot act under their own volition, of course). But many generally successful, hard-working, intelligent, learners in chemistry classes all over the world seem to think otherwise (Taber, 1998 [Download paper]).

Read about the octet framework – an alternative conceptual framework

Likewise, electrons do not care if they are in an atom or not, or whether they are spin-paired or not (and if so, which other, indistinguishable, electron they are paired with), or which energy level of a system they populate.


header from published paper

The authors of the recent paper (which is open access, so freely available for anyone who wishes to download/read it) claim that students found the story-related activity engaging (which certainly seems likely) and that it helped address some misconceptions about atomic structure. They note that:

  • "Students do not clearly understand the concept of an orbital" (Aquilina, Dello Iacono, Gabelli, Picariello, Scettri & Termini, 2024)

This is a topic that has long interested me so I took a look at the activity the researchers had devised. The learners were

"10th-grade classes, with the participants' average age being between 15 and 16, attending a technical computer science high school 1…[who] had already studied the atomic model in their chemistry classes during the first half of the year."

Aquilina, Dello Iacono, Gabelli, Picariello, Scettri & Termini, 2024

I have taught a basic (planetary) model of atomic structure to students at this age, and also more advanced models to 16-19 year old learners (on A level courses), so I was keen to read about the activity. The authors did not include an explicit statement of the curriculum content which was being treated as target knowledge, although they did include a discussion of their rationale for the story, as well as comments on student work, from which some features could be deduced or inferred. (I would have found it useful to have read an explicit statement of just what the learners were expected to know – what the 'correct' model was meant to be – at the outset of the paper.)

I approached the paper thinking it was ambitious to teach an orbital model of the atom to students of this age. My reading of the story (reproduced below) reinforced that initial impression (I admit, I was challenged in places!) – although the authors certainly felt the students in their research coped well with the challenge.

Although I felt I struggled interpreting some features of the narrative,

A student with a specific learning disorder (SLD), mentioned, "The connection of a fairly complicated topic with such a simple story"

Aquilina, Dello Iacono, Gabelli, Picariello, Scettri & Termini, 2024

It is important to note that the teaching scheme adopted a dialogic approach, where class discussions were included at two points after the students had worked in groups on parts of the activity. The activity was also conceptualised as being part of an enquiry-based learning cycle. So, the material below should be read accordingly, as it does not reflect this wider classroom context.

Read about dialogic teaching

Read about enquiry-based science education


The story

The story is broken into four parts, each leading to a task for the learners (working in groups) to engage in.


Prologue

"Romeo is a bold and dynamic electron found in an atom with seven energy levels. He is at the 4s energy level, together with the faithful Mercutio, his companion on raids. Always upside down compared to him, but then there is no place for two equal electrons in their crew. The two are part of the Montague family, known for being particularly lively.

Juliet is an electron in 2s, she is more tied to her nucleus and in fact she is a Capulet, a rival family to that of the Montagues and decidedly more calm. Juliet is always accompanied by her nurse; they too are turned upside down with respect to each other.

There is a grand ball to which everyone is invited, and, to better organize their arrangement, there is a need to schematize their position."

[Instructions to learners: "Discuss with your classmates what should be the design of the atom where the two families «are» and build
a model"]

Aquilina, Dello Iacono, Gabelli, Picariello, Scettri & Termini, 2024

Chapter 1 – part 1

"At one point during the dance, Romeo notices Juliet in her orbital, and, even if he occasionally gets close to her, he is unable to stay there permanently: quivering with love, he asks who knows her and what her tastes are in terms of radiations (electrons are well known to be romantics). He discovers that Juliet is obsessed with color harmony and that the color she prefers is purple "486 nm". To get noticed he wants to perform his famous photon–spectroscopic serenade and jump to emit a purple trail.

[Instructions to learners: "Discuss with your teammates to help Romeo understand how far he will have to jump and whether or not he would have gotten closer to Juliet in this way."]

Aquilina, Dello Iacono, Gabelli, Picariello, Scettri & Termini, 2024

Chapter 1 – part 2

"The two are deeply in love and would like to spend the rest of their days together. But Juliet's family hinders them, crying scandal: a Montague cannot be so tied to the nucleus! What to do? The nurse offers Romeo the chance to take her place, but, for her, this would mean losing her place next to Juliet. Romeo and Juliet, very hesitant, then decide to move towards the orbitals occupied by the Montagues. But how to get up there?

While the couple is tormented by this problem, an enlightened friar, Lory, arrives to their rescue with two THz 457s, offering to give them a lift. Despite this help, Romeo and Juliet are unable to reach the Montague orbital, so they loudly invoke another friar, Enzo, asking for new help.

[Instructions to learners: Discuss with your teammates to understand how far they will jump thanks to the first photons and which photons Fra Enzo will have to carry for the two lovers to reach the Montague orbital."]

Aquilina, Dello Iacono, Gabelli, Picariello, Scettri & Termini, 2024

Chapter 2 and epilogue

"Juliet's escape has thrown the entire atomic balance into crisis, forcing some Montagues to change levels in order to maintain overall stability. Then, when the couple comes to the Montagues, they cry out for revenge, and the couple is then forced to flee again.

The Montagues set out in search of Romeo and Juliet but fail because it is not possible to reconstruct the trajectory followed by the two lovers.

The story unfortunately ends in tragedy: the two do manage to free themselves from the influence of their families, but they still understand that they cannot be together. Now condemned to separation, the two lovers decide to draw up a schema of the place (the atom) where they met to remember it forever.

[Instructions to learners: "Discuss with your teammates why this trajectory cannot be reconstructed. End the story with a tragic ending, explaining the reasons for the separation sentence.

EPILOGUE Construct with your teammates a possible model of the scheme realized by Romeo and Juliet."]

Aquilina, Dello Iacono, Gabelli, Picariello, Scettri & Termini, 2024

Interpreting the narrative

Reading the account I had a very mixed response. I am very keen on approaches that use the familiar everyday as ways into teaching complex, abstract ideas; but subject to two provisos:

  • these everyday analogies are interim supports ('scaffolds'), to be withdraw as soon as they are no longer needed;
  • teaching needs to focus on the 'negative analogy' (things that do not map across) as well as the 'positive analogy' (the aspects of the comparison that 'work').

The approach here seemed somewhat different. The learners had already been taught a model of the atom earlier in the year, and this activity was intended to be an opportunity to review this prior learning and apply it – and an opportunity for teachers to identify any alternative conceptions elicited by the activity.

Metaphorical meanings?

Romeo and Juliet are not the lovers in the stage play, but electrons. Therefore, in reading the story I identified scientific information (electron Romeo is in a 4s orbital in an atom) and material that seemed to be metaphorical (the electrons Romeo and Mercutio go on 'raids'). I therefore saw the task of reading the story as being in part a decoding of the metaphors that were used.

So, the idea of Romeo and Mercutio being relatively "upside down" was not to be taken literally (electrons do not have ups or downs) but to be a metaphor for spin +1/2 and spin –1/2, often referred to metaphorically as 'spin up' and 'spin down'. Going on raids was more tricky: in some chemical reactions electron pairs are considered to shift during bond formation (or bond breaking, but that would not refer to an atomic species), but 'raid' suggests a temporary excursion.

I could not understand in what sense Mercutio (the electron, not the fictional character) could be said to be faithful. Electrons respond to physical forces, not personal attachments. Perhaps, I was over-thinking this, and not all the narrative elements did map onto the atomic system? Perhaps that was meant to be part of the challenge for the learners?

A fundamental concern with this kind of comparison is that all electrons are inherently identical, and are only distinguished by the accidental features they acquire in a particular system.

  • A 2s electron is on average closer to the nucleus, and experiences a greater effective core charge (it is not shielded as much from the nucleus as a 4s electron is) – so the 'tie' (bond) to the nucleus can be understood as analogous to the attractive force operating between the electron and nucleus. 2
  • The reference to being more calm perhaps refers to how the 2s level is at a 'lower' energy so the 'particularly lively' 4s electrons can be more dynamic?

If Romeo and Mercutio, or even Romeo and Juliet, were swapped it could make absolutely no difference and no one could tell. By giving electrons personal identities they seem to be more like us and less like electrons. Electrons cannot be bold or calm. Romeo and Juliet behave differently because they are in different orbitals at different energy levels, not because they are different electrons. Could learners miss this critical point? If Juliet (or Romeo) moved to a different energy level then she (or he) would change 'personality' – but that would undermine the narrative.

I was not sure how the two families related to anything. Within an atom we could class some electrons alike because they are in the same 'shell' (have the same principal quantum number) – so perhaps the two families were in the n=2 and n=4 levels (the L and N shells being their metaphorical 'houses'). I also could not understand where the ball was meant to be held:

  • were the electrons to be moved to a new set of orbitals (requiring promotion)
  • were the electrons meant be moved to outside the atom (requiring ionisation), or
  • was the ball to take place with the electrons in their current orbitals (but for some reason behaving differently than when no dance was taking place?)

The attraction between Romeo and Juliet (the electrons, not the fictional lovers) was difficult to understand. Certainly, if we adopt a model of electrons moving about in different orbitals 3 then they could sometimes be nearer to each other as atomic orbitals interpenetrate – and if so they would influence each other more (due to their charge and spin) at these times: but this would primarily be a repulsion.


Interpenetrating fields of play. If two sports pitches were marked out overlapping on the same ground, then there would be places that were part of both fields of play.

(Consider a school with very limited space for sports pitches. Perhaps they mark up a soccer pitch and a field hockey pitch overlapping. If both soccer and hockey players train at the same time there will be places that are part of both pitches, and players from the two sports can come close together in those areas. {This is just an analogy. The two sports would need to schedule practice at different times to avoid accidents!})


It seemed to me that the learners were being asked to read the account at two levels – some features of the story were metaphors (such as when the lovers left the atom only to find they had separate indeterminate trajectories) when other features seemed to be simply plot devices to provde an engaging narrative. I thought that the students were being asked to work out which bits of the story they should take seriously as corresponding to part of an atomic model, and which just moved the narrative on. I though this might be challenging for the 14-15 year old learners (as I was struggling!)

Orbitals and transitions

Some features of the story seemed potentially likely to encourage alternative conceptions. Juliet's preference for light of wavelength 486 nm risks the association of a spectral line with an electron or an energy level, rather than with a transition.

The specific references to 486 nm and 457 THz radiation seemed to suggest that a quantative model was needed – where an atom would actually show spectral lines reflecting transitions associated with radiation of these specific characteristics.

The rationale

Unlike the students, I had access to some of the resource designers' thinking as the paper included a rationale for the storyline. This acknowledged that

The specific location of the grand ball remains implicit [?], as it is challenging to conceive of electrons dancing outside the metaphorical context of "moving swiftly". However, all the other character details are essential for initiating the story and allowing mathematical and physical problems and situations to emerge."

Aquilina, Dello Iacono, Gabelli, Picariello, Scettri & Termini, 2024

This seemed to confirm that the learners were expected to build a quantitative model. This was reiterated later in the rationale

"Through calculations of energy transitions and the resulting orbital distances, students gain insight into the quadratic proportionality that underlies these phenomena [?], prompting a gradual reshaping of their personal notions regarding orbital distances."

Aquilina, Dello Iacono, Gabelli, Picariello, Scettri & Termini, 2024

I was not sure what was mant by 'orbital distances', and return to this point below. I was also not sure how quadratic proportionality underlay energy transitions.

This was only one of the points in the paper where I got the impression that in the teaching model adopted, energy levels and orbitals were not only being associated, but at times almost seen as equivalent and interchangeable.

A diagnostic assessment opportunity

The rationale seemed to confirm that the activity was deliberately testing whether students associated spectral lines with energy levels rather than transitons between levels,

"To elucidate the intriguing connection between emission and electron transitions to different energy levels, we introduce a romantic-comedic twist, employing Juliet's passion for color harmony as a plot device. Juliet's preference for the color purple is strategically chosen to align with her energy level, prompting students to contemplate the intriguing relationship between spectroscopy lines and electron energy transitions."

Aquilina, Dello Iacono, Gabelli, Picariello, Scettri & Termini, 2024

On the other hand, my suspicion that I had been reading too much into the narrative, and trying too hard to interpret plot twists was rather undermined by being told,

"Take, for instance, Romeo's desire to gain Juliet's attention and their joint pursuit of a life away from their feuding families. This narrative intricately parallels the fundamental interplay of orbitals within the model, establishing a direct and compelling link between the characters' human drama and the pivotal role of orbitals in the model."

Aquilina, Dello Iacono, Gabelli, Picariello, Scettri & Termini, 2024

Indeed? I was struggling to map across some of the story, even when (unlike the students) I had access to the rationale:

"At the outset, the consequences of Romeo and Juliet's choices become apparent: the voids within the nucleus [?] are replenished with new electrons [?], ultimately disturbing the equilibrium of the two feuding families. This disruption leads them to share orbits [sic], not fueled by anger but by fate. The Montagues seek revenge, yet they grapple with the inability to reconstruct the electrons' orbitals due to the uncertainty principle."

Aquilina, Dello Iacono, Gabelli, Picariello, Scettri & Termini, 2024

A lot of this went over my head.

The uncertainty principle would not interfere with characterising orbitals, only with being able to posit specific electron trajectories. The orbitals do not belong to electrons ("the electrons' orbitals") but are characteristic of an atomic system with its configuration of charges.

A hybrid model?

Perhaps, in part, my confusion was due to my not being clear about what the target knowledge was- exactly which kind of model was it hoped the students would produce?

"After studying the planetary and Bohr atomic models, students cannot easily move beyond them"

Aquilina, Dello Iacono, Gabelli, Picariello, Scettri & Termini, 2024

It seemed clear from the paper that the learners were expected to have moved beyond a model with planetary orbits, to a model with orbitals, and so from the idea of electrons moving on definite trajectories, to being found somewhere within the orbitals. 3

There was historically a range of models of the atom (even 'the Bohr model' was actaully a series of models), and long ago Rosaria Justi and John Gilbert (Justi & Gilbert, 2000) pointed out that often in teaching we end up presenting 'hybrid' models – that is, models which have features drawn from across several of the different scientific models. Did the curriculum these students followed set out such a hybrid model for students to learn? 4

An atom with seven energy levels?

At the start of the story, the students were told "Romeo is found in an atom with seven energy levels". I am not sure any real atom could only have seven energy levels. My understanding is that any atom has in principle an infinite number of energy levels, but the the spacing of the levels gets successively smaller, so they converge on a limit (which makes ionisation feasible). Even the hydrogen atom has an infinite number of energy levels, but only one is populated with an electron.

So, I wondered if possibly this was meant to be read as "Romeo is found in an atom with seven populated energy levels"?

A sensible starting point for a student is to assume the atom is initially in its ground state (as under normal circumstances they usually are). If the reference to seven energy levels means populated energy levels, and students are to assume the atom starts in the ground state then presumably learners are meant to assume the atom they need to model is one of the first transition series (i.e., elements with electronic configurations from 1s2 2s2 2p6 3s2 3p6 4s2 3d1 to 1s2 2s2 2p6 3s2 3p6 4s2 3d10: that is an atom from one of the elements scandium to zinc).

However, later there is a reference to electron Romeo wanting to "jump to emit a purple trail". But he needs to jump 'down' (to a lower energy level) both to get closer to Juliet and indeed to "emit a purple trail" (i.e., for Romeo to be promoted, light would need to be absorbed not emitted) – which is only possible if the atom is NOT initially in its ground state, so that there will be an orbital at a lower energy level not fully occupied. That potentially complicates the model to be built.

For one thing, if the atom is not in its ground state, then atoms of elements of lower atomic mass than scandium might be the target atom to be modelled? Indeed, any atom from the element nitrogen (in the highly excited configuration 1s1 2s1 2p1 3s1 3p1 4s1 3d1 ) on to zinc could theoretically have seven occupied energy levels. It did not help that there seemed to be no information on how many electrons were in this atom – four were specified, and we are told unspecified other 'family' members lived there, and two other characters were name-checked without it being explicit if they were also in the atom or just passing (from the local Abbey perhaps – would that be an atom of a noble gas?)

Interorbital distances?

As noted above, the authors refer to how they "delve into the concept of interatomic orbital distances", but this seems an oxymoron.

"From the analysis of the drawings, it emerges that the students' final drawings can be traced back to three different types of atom representation (R):

  • R1: orbits/orbitals represented at varying distances to convey the concept of energy levels more effectively;
  • R2: orbits/orbitals represented at correct distances according to the radius;
  • R3: attempt to depict the concept of orbitals and the correct distances between them."
Aquilina, Dello Iacono, Gabelli, Picariello, Scettri & Termini, 2024

The authors refer to how in a figure assigned to category R3, "The distances between the spheres reflect the correct distances according to n2", but this does not strictly relate to an orbital model.

Orbitals do not have edges, so it is not possible to measure how far they are from anything. Strictly, every orbital reaches to infinity (even if the electron density soon gets so rare that it becomes effectively zero). The point is that this is a gradual falling-off and there is no sudden drop that we might think of as an edge.

Commonly orbitals are represented either with

  • probability contour lines, or
  • colour or shading showing differnt levels of electron density (i.e., the relative probabilities of an electron in the orbital being 'found' at different regions of the orbital), or
  • more simply with probability envelopes.

Those envelopes show where, say, 90% or 95% of the electron density is located – which means 10% or 5% of the electron density (that is inside the orbital) lies outside the envelope drawn. So, these lines are to soem degree arbitrary, conventional and do not correspond to anything physical ('real').

One could measure the distance between the centres of two different orbitals, but this would be a trivial issue when the orbitals are in the same atom. (That is, the atomic orbitals are all centred on the nucleus, so the centres have no distance between each other.)

This is different to a planetary type model where electrons are considered to be a certain distance from the nucleus, so the orbits have quantifiable radii. In moving to an orbital model we have to think of fuzzy overlapping volumes of space, and the notion of there being set distances between orbitals just does not work in this model.


Imagine being asked to report the distance between the soccer pitch and the hockey pitch.


And then imagine having that task when there are no marked out edges to the pitches.


The energy levels associated with the orbitals can be considered to have specific values, and so there are definite differences ('distances'?) between the levels in that sense – but these would be energy gaps: analogical 'distances' on an energy scale, not actual distances.

The authors suggest that,

Despite their discussion about orbitals, [for the students' final drawings] all groups drew orbits, representing them as lines depicting the trajectories of electrons

Aquilina, Dello Iacono, Gabelli, Picariello, Scettri & Termini, 2024

But that is not so clear from the diagrams of the models and the students' own comments.

Student 1: "In a circle, we drew lines. But we know that electrons don't follow that precise path; they exist in orbitals, which are regions where electrons are more likely to be found. So, we don't know the precise radius because it's a region. Therefore, in my opinion, since the radius can always vary, you can't use the radius to depict the atomic model; it's more accurate to use energy levels."

Teacher: "Here you have drawn the distances increasingly closer. Why?"

Student 2: "Because it represented differences in energy levels."

Aquilina, Dello Iacono, Gabelli, Picariello, Scettri & Termini, 2024

Some groups of students seem to have drawn concentric circles representing energy levels rather than orbits or shells or orbitals. Normally, energy level diagrams are not drawn like that, but this seems a perfectly reasonable form of representation providing it is explained.

Spherical orbitals

We also have to bear in mind that only s-orbitals have spherical symmetry. (A 'shell' of orbitals in an atom would be spherically symmetrical only if each orbital was singly or fully occupied. But it was not clear how many electrons were in this atom.)

The first seven energy levels in any atom or ion with more than one electron will be associated with p- and d-orbitals as well as s-orbitals. So, even if orbitals were represented with probability envelopes, and these were treated (incorrectly) as if the edges of the orbitals, then there would be no fixed 'distances' between the edges of any comparisons involving these non-spherical orbitals.


image of orbitals

Not all orbitals have spherical geometry (Image by Smiley _p0p from Pixabay)


At this point it is interesting to examine the samples of student models represented in the paper. All of them are drawn with circles. The authors of the paper seemed satisfied with this aspect of the models.

Making sense of 486 nm and the 'THz 457s'

I pointed out above that my reading of the information given about the atom that it seemed the target atom could be from one of a wide range of elements. It seems I got this completely wrong,

We conclude this paper by highlighting a limitation of the story we have designed from a physical point of view. Our story does not fit the real atomic structure. Indeed, we chose to consider a hydrogen atom with multiple electrons because we thought it was easier for the students to manipulate. We are aware of the fact that this may represent a critical point of our story, but in the classes where we experienced the activity it has not created problems, since the students noticed this inconsistency and talked about it with the teacher.

Aquilina, Dello Iacono, Gabelli, Picariello, Scettri & Termini, 2024

Now, by definition, a model is never quite like what is modelled – or it ceases to be a model and becomes a perfect replica. But "a hydrogen atom with multiple electrons" is not an atom at all, but an ion. I am not clear why this is "easier to manipulate" than an atom of a different element, as in models of this kind the nucleus is in effect just a minute point charge – so its composition does not complicate the model in any significant way. If that nuclear charge is +7, say, rather than +1, it makes a difference, certainly (to energy levels), but that does not add any further complexity.

Perhaps the authors chose to retain a hydrogen nucleus because they wanted students to use data from hydrogen spectra? (But if so, this was a little naughty.)

The Balmer series

Again, it did not help that I did not know what the target knowledge set out in the curriculum was.4 But, knowing now that hydrogen was the target atom led me to suspect 486 nm and 457 THz radiation linked to lines in the hydrogen spectra – lines in the Balmer series associated with transitions between n=3 and n=2 (656 nm) and n=4 and n=2 (486 nm).

That was all very well, but those transitions referred to the hydogen atom and not to a hydrogen ion. The extra electrons repelling each other in the ion (assuming the ion could be considered stable, which is itself problematic) mean the energy levels (and so the energy gaps; and so the spectral lines) would all be different.

But, if we pretended the ion was stable, and if we pretended that the additional electrons did not change the energy levels (what is what I meant by being somewhat naughty), then the numbers made sense.

A sleight of hand?

Indeed, if we were to adopt the hydrogen atom as the model for our ion, then I sensed I understood why the orbitals were all drawn as circles. In the hydrogen atom, the energy levels are only associated with the principle quantum number. The 2p orbital is at just the same energy level as the 2s orbital. A transition from the N shell to the L shell has the same energy associated with, and so the same frequency of radiation, regardless of whether it involved 2s-4s or 2p-4s or 2s-4p or 2p-4p or 2s-4d or 2p-4d (or indeed 2s-4f or 2p-4f)5. That is a considerable simplification, that would make the task much easier for learners.

So, if we are modelling the hydrogen atomic energy levels, we only need to worry about the principle quantum number as there is one level for each value of n. The student diagrams reproduced in the paper suggested all the students understood the reference to an atom with seven energy levels to mean n (that is the principle quantum number related to 'shell') = 1-7.

But an energy level is not an orbital. The n=2 energy level in a hydrogen atom is associated with 4 orbitals, only one of which has spherical symmetry. The n=3 level is associated with 9 orbitals, only one of which has spherical symmetry.

Moreover, this assumption that all the orbtials in a shall are at the same energy level ('degenerate') only applies to a hydrogenic species (H, He+, Li2+, etc.) – that is, atom-like species with a single electron. The 'atom' (ion) with Romeo and Juliet and Mercutio and the nurse and the rest of the Capulets and Montagues (and possibly some clergy) would not have 2s and 2p orbitals that were degenerate. The presence of interacting electrons (repelling each other, that is, not lusting after each other and "quivering with love") would raze the degeneracy- so the 2s and 2p orbitals would actually be at different energy levels. And so also with 3s and 3p and 3d.

It is not the presence of a hydrogen nucleus which leads to degeneracy between the orbitals within each value of n (each shell), but a system of one nucleus and one electron. So if this 'atom' (ion) had seven energy levels, these would not equate to seven shells of electrons.

The model

So, it looks like the target model was an ion with a hydrogen nucleus, and 7 energy levels occupied by an unspecified number (>4) of electrons, which has the same energy structure and levels as a hydrogen atom, but where each energy level only contained an s orbital.

Models simplify, and in modelling we deliberately leave aside some complexity and nuance. However, we have to balance the gain in simplicity with the loss of authenticity.

  • A highly charged hydrogen ion could not exist (unless maintained by some very powerful external field)
  • Atoms have an infinite number of energy levels (but there is no harm in asking learners to ignore most of them for the time being when working on a task)
  • A hydrogen atom has orbitals of different types (s, p, d…) not all of which are of spherically symmetrical.
  • The electronic transitions in an ion would not be those found in the related atom, as energy levels of the system depend on the configuration of charges that are interacting. The ion would have many more potential transitions than a single-electron system (such as a hydrogen atom), and these would not have the same energies/frequencies/wavelengths as in the hydrogen atom.
  • Orbitals do not have edges, and they interpenetrate, so the concept of interatomic orbital distances does not correspond to anything 'realistic' in the orbital model of the atom.

So, the model seems to put aside a lot of the subtlety of the science. But then are these nuanced ideas suitable for treatment with most 15-16 year olds? I would have suspected not (which is why I started from a position of thinking this whole activity was somewhat ambitious), and that may well be why compromises were made in the teaching model adopted in this study.

But perhaps it would be better not to introduce an orbital model until we feel learners are ready to appreciate the quantum jump from concentric orbits to fuzzy, overlapping, infinitely-extended patterns of electronic probability, and the associated complex patterns of energy levels they generate. (But, again, the teaching model used may simply have been reflecting the target knowledge set out in the school curriculum in this particular national context? 4)

After all, as the authors had noted,

"Students do not clearly understand the concept of an orbital" (Aquilina, Dello Iacono, Gabelli, Picariello, Scettri & Termini, 2024)

Encouraging a new alternative conception?

To take one point. The 486 nm and 457 THz radiation is associated with transitions between n=3 and n=2 (656 nm) and n=4 and n=2 (486 nm) in the hydrogen atom, but NOT in the 'atom' populated with Montagues and Capulets.

Does this matter? After all, the point of the exercise is not to remember these specific values, but to be able to link radiation emitted or absorbed to electronic transitions – so, the particular values of 486 nm and 457 THz are irrelevant. True, but what students are potentially learning here is that the values of energy levels are not affected by the number of electrons repelling each other (here we have an ion with many electrons, but we can simply use the values for a hydrogen atom) – which is an alternative conception.

I also know that this is an alternative conception that learners are likely to readily develop. When students study ionisation energies, and make comparisons between different atoms, they often fail to allow for how the same designation of orbital does not imply an equivalence between differently populated electronic structures.

So, for example, a 2p orbital in an oxygen atom is not only not equivalent to a 2s orbital in the same atom: nor is it equivalent to a 2p orbital in a nitrogen atom. Nor, for that matter, is it entirely equivalent to a 2p orbital in the o2- anion.

This is not the most serious alternative conception that students can acquire, but given the complexity and challenge of this whole topic area, it might be wise to avoid risk misleading students when possible.

Or am I just being over-critical because I myself found the task too challenging? ☹️

To see through an orbital clearly?

This was an interesting project, and I hope the authors explore the idea further, and perhaps use their experiences with this implementation to further refine the activity. But I am not sure it is helpful in the long term to encourage learners to work with a model that is so constrained that it is likely to encourage new alternative conceptions.

But would that be the case? If the activity is part of a dialogic teaching sequence and the catalyst for engaging students in a discussion of these abstract ideas – a discussion that the teacher carefully steers towards the canonical account – then perhaps the outcome can be more productive. I guess we can only conjecture about this, until someone investigates the long-term effects of learning from the activity.

As usual, it is fair to say "more research is needed".



Work cited:

Aquilina, G.; Dello Iacono, U.; Gabelli, L.; Picariello, L.; Scettri, G.; Termini, G. "Romeo and Juliet: A Love out of the Shell": Using Storytelling to Address Students' Misconceptions and Promote Modeling Competencies in Science. Education Sciences, 2024, 14, 239. https://doi.org/10.3390/educsci14030239

Justi, R., & Gilbert, J. K. (2000). History and philosophy of science through models: some challenges in the case of 'the atom'. International Journal of Science Education, 22(9), 993-1009.

Taber, K. S. (1998) An alternative conceptual framework from chemistry education, International Journal of Science Education, 20 (5), pp.597-608.
[Download paper]

Taber, K. S. (2002) Conceptualizing quanta – illuminating the ground state of student understanding of atomic orbitalsChemistry Education: Research and Practice in Europe, 3 (2), pp.145-158 [Download paper]

Taber, K. S. (2019). The Nature of the Chemical Concept: Constructing chemical knowledge in teaching and learning. Royal Society of Chemistry.

Taber, K. S. and Watts, M. (1996) The secret life of the chemical bond: students' anthropomorphic and animistic references to bondingInternational Journal of Science Education, 18 (5), pp.557-568. [Downlod paper]


Notes

1 Of course there are many atoms, and indeed many kinds of atoms – so the use of the definite article ('the') is strictly inappropriate. But, this is common usage,

What seems potentially more problematic is the use of the definitive article when the referent is not a specific individual specimen. Chemistry teachers will say things like "the ammonia molecule is pyramidal" when no ammonia molecule is either specified directly or can be inferred to be the case in point from the context. This probably does not seem problematic for the simple reason that it does not matter which ammonia molecule is being referred to: they are all pyramidal. So, statements such as the ammonia molecular is pyramidal; the chlorine atom readily accepts an electron; the K shell is nearest the nucleus; and the iodide ion is a good leaving group; etcetera, will be true regardless.

These statements 'work' in a way that some apparently parallel statements from outside of chemistry would not: the house has a blue door, the man walks with a limp, the baby sneezed all night, the bicycle has squeaky brakes, etcetera. Some houses have blue doors – many do not…So, we should not say 'the house has a blue door' unless we have made it clear which house we are referring to. Yet, we do not need to say which particular water molecule is polar, as they all are (i.e., it may be considered an essential quality of a water molecule). So, the question here is why a teacher would say 'the ammonia molecule is pyramidal' when they are not actually referring to a particular specimen, and the point they are making is actually that (all) ammonia molecules are pyramidal.

Taber, 2019, p.128

And, even if we can refer to 'the carbon atom' when we mean any and all carbon atoms, to simply refer to 'the atom' seems a slight to the periodic table – surely we need to say which (kind of) atom we are modelling? That point certainly proved to be critical in the context of the modelling task discussed in this article!


2 The force is symmetrical – the same magnitude force acts on the nucleus and the electron, with each being pulled towards the other. Students commonly have alternative conceptions about this such as thinking the force only acts in one direction (from nucleus to electron) or that the force on the electron is greater.

Read about Newton's third law and common alternative conceptions


3 In the planetary model of the atoms, electrons moved in orbits. In the orbital model we can think of electrons moving about the orbital, and the 'electron density' as a kind of average over time of where they have been. However, it may be more in keeping with the quantum model of the atom to suggest the electrons do not actually move around but rather have probabilities of being located at different points under conditions of observation. (According to a very common interpretation of quantum theory, the notion of an electron being somewhere specific only makes sense at the point of observation.) This is pretty difficult to appreciate (especially for most school-age learners), and I suspect most chemists are happy enough most of the time to think of the electrons moving around in their orbitals.


4 Five of the six authors, including the corresponding author, were based in Italy (the other author gave an affiliation based in Canada), so I assume the schools from which the work is reported is in Italy. The paper reports the task set and the student responses in English, so it is not clear if English was used as the language of instruction in the school (this seems unlikely unless this was an International School, but the paper does not report that material has been translated into English).


5 4f orbitals are not usually relevant to atomic structure till we consider cerium, element 58. But the familiar order of filling orbitals as we imagine we are building up atoms (1s < 2s < 2p< 3s < 3p < 4s < 3d < 4p… *) refers to species with more than one electron. For a hydrogen atom, a 4f orbtial is at the same energy level as the 4s orbital, as when occupied the atom's electron, neither would be sheilded at all from the nucleus by other electrons.

(* Ironically, the familiar descriptions of the discrete orbitals designated in this way are based on calculations for a hydrogen atom and do not strictly apply to multi-electron atoms. However the moodel generally works well, and is widely used.)


Surface tension is due to everybody trying to get into the water

Surely you are joking, Prof. Feynman? 1


Keith S. Taber


Photo of Richard Feynman, taken in 1984 © Tamiko Thiel (accessed from Wikipedia and shared under Creative Commons Attribution-Share Alike 3.0 Unported)


The late, great, Richard Feynman

Richard Feynman was special. Any one who wins the Nobel prize has to be pretty special, but physics laureate Feynman was even more remarkable as he was an exceptionally high achieving research physicist also known for his…teaching. No one gets a Nobel for being a good teacher, and it is often considered in Academia that teaching (that is, if one tries to give teaching the time and energy required to do it well – as students deserve) distracts from research to such an extent that it is rare to excel in both.

Feynman had something a lot of scientists do not not: great charisma. (That is no insult to fellow scientists – most plumbers and greengrocers and bus drivers and accountants and hairdressers do not – that is what makes it notable). He might be considered the Albert Einstein of the second half of the twentieth century, and because of that timescale we are lucky to have quality recordings of him talking and teaching in a way that could not have happened with previous generations. (A great shame in many cases: if perhaps a blessing with some – Isaac Newton's lectures were apparently avoided by most of his own students.)

Like many people, I find Feynman bewitching – he had a sparkle about him – almost a permanent mischievous twinkle in the eye – and an ability to somehow express the excitement of science (of working out why things are as they are) whilst being able to talk in ways that could be understood by people that lacked his expertise. That is perhaps one trait of a great teacher – being able to talk at the level of the audience, despite personally understanding at a higher, more complex and subtle, level.

That is by way of preamble – as I want to consider an explanation Feynman once offered of surface tension.


Screenshot of Richard Feynman explaining why water forms into drops.


Why does it rain in drops?

The extract I am discussing is taken from a 1983 BBC series of short episodes in a series called 'Fun to Imagine'. Although, at the time of writing, the episodes are "not currently available" from the BBC site, there is a compilation on YouTube. One of the topics Feynman discusses is the origin of surface tension – although he only introduces the technical term after explaining the phenomenon that water forms into droplets,

"you see a little drop of water, a tiny drop
And the atoms [sic, molecules] attract each other, they like to be next to each other
They want as many partners as they can get
Now the guys that are at the surface have only partners on one side
here, in the air on the other side, so they're trying to get in
And you can imagine…this teeming people, all moving very fast
all trying to have as many partners as possible and the guys at the edge are very unhappy and nervous and they keep pounding in
trying to get in, and that makes it a tight ball instead of a flat
and that's what, you know, surface tension
When you realise when you see how sometimes a water drop sits like this on a table then you start to imagine why it's like that
because everybody is trying to get into the water"

Richard Feynman speaking in 1983

Is this a good explanation?

Well, we might suggest Feynman makes a schoolchild error – water is not an atomic substance, but molecular. It does not contain discrete atoms, so he should be referring to the molecules attracting each other. But I do not think this is an error in the sense that Feynman was mistaken, simply that although the distinction is of great importance in chemistry, physicists sometimes use the term 'atom' generically to refer to the individual particles in a gas, for example. That might be unhelpful to a secondary school student studying for examinations, but if Feynman thought of his television audience for the recording as lay people, the general public, then perhaps the distinction between atoms (arguably a more familiar term in everyday discourse) and molecules would be considered an unhelpful detail? I am certainly prepared to give him that. I think it was the wrong choice, but not that Feynman was in error.

But what about the overall argument here. The 'atoms' want to have partners all around them 2 so they try to get inside the volume of the liquid. The overall effect of everyone, including these guys at the edge, trying to get inside the water is that it forms a sphere-like shape: "a tight ball instead of [something more] flat". Is that a convincing explanation – and is it a valid one?

What makes for a good explanation?

If anything is central to both science and science teaching, it is explanation.

"Explanation would seem to be central to the essence of science. A naïve view might claim that science discovers knowledge about the World, although it might be more accurate to suggest that science creates knowledge through the development of theories. The theories are used in turn to understand, predict and sometimes control the world, and in these activities, scientific explanations play the key role. We might consider theories and models to be the resources of science, but explanations to be the active processes through which theory is applied to contexts of interest…

An explanation is an answer to a 'why' question: but that in itself neither makes for a good explanation, nor for a scientific one. There is no simple answer to what does count as a good explanation, in science or elsewhere. Explanations have audiences, and to some extent, a good explanation is one that satisfied its audience – in other words it meets the explainee's purpose in seeking an explanation. Additionally, it has been known since at least Aristotle's time that we can talk of different kinds of causes, which suggests that many 'why questions' might have different types of acceptable responses, depending on the type of cause being sought."

Taber, 2007, p.159 [Download the chapter]

That passage is taken from a chapter where I described some activities used with secondary school students to help teach them about the nature of scientific explanation. (Read about the classroom activities here.) In that context, working with learners who were about 14 years of age, students were told that a good scientific explanation would be logical, and would draw upon scientific theory,

"pupils were told that scientific explanations needed to take into account logic and theory, i.e., that the explanation needs to be rational, and the explanation needs to draw upon accepted scientific ideas. As the notion of 'theory' is itself known to be difficult for students, they were also told that scientific theories are ideas about the world which are well supported by evidence; are internally consistent; and which usually fit with other accepted theories."

Taber, 2007, p.159 [Download the chapter]

Feynman's explanation is logical (if incomplete)

In that regard, Feynman's explanation can be considered logical, even if it omits (i.e., he takes as assumed) an important step* that is needed to explain the (approximately) spherical shape of the water drop.

If water quanticles (let's leave aside whether they are atoms or molecules) want to have many partners 2, and so try to get inside the volume, then we can understand* that the water drop will tend to the smallest surface area possible, so as few quanticles end up at the surface (with the tenuous air, rather than congregating water partners, on one side) where they will be nervous, and as many quanticles as possible are in the interior of the drop where they will be happy.

* The missing step is to state that a spherical drop will have a smaller surface area than any other shape with the same volume and so fewest quanticles at the surface. Perhaps Feynman assumed everyone would know/see that. Probably there is no such thing as a totally complete explanation.

So, is this a good explanation?

Explanations can have different purposes. Scientific explanations allow us to make effective predictions (and so often to control situations – the application of science through technology). But, in everyday life, explanations have a more subjective purpose ("explanations have audiences, and to some extent, a good explanation is one that satisfied its audience").

If, as a result of hearing Feynman's explanation, the viewers of the BBC televison programme

  • felt they now understood why sphere-like drops of water form, and
  • considered they had made sense of some science, and so
  • appreciated the value of science in explaining everyday phenomena,

then perhaps the explanation achieved its purpose?

Was Feynman's explanation scientific?

Of course, if I am being my usual pedantic self, I could point out that although Feynman's explanation was logical, that does not make it scientific unless it also drew upon accepted scientific principles. It was logical because the explicandum (what was to be explained – here, the drop shape) followed from the premise (i.e., if water quanticles want to have many partners, and act accordingly, then…)

But, in science, quanticles are not understood as sentient actors, but as inanimate entities that are not (and cannot be) aware of their situation and cannot act deliberately to work towards personal goals. Therefore, no matter how convincing someone may have found this explanation, it does not qualify as a scientific explanation as it is not based on accepted scientific principles (…or at least, not directly).

An anthropomorphic explanation

Feynman's explanation uses anthropomorphism, which from a scientific perspective makes it a pseudo-explanation. A pseudo-explanation takes the form of an explanation in that it is presented as if an answer to a why question, but does meet the requirements for a formal explanation (e.g., it does "not logically fit the phenomenon to be explained into a wider conceptual scheme", Taber & Watts, 2000.)

There are various kinds of pseudo-explanations such as tautology (circular explanations that rely on the conclusions as premises) and simply offering a label for the explicandum (e.g., water absorbs a lot of heat for a small change in its temperature because it has a high heat capacity – this is a kind of disguised tautology, as a 'high heat capacity' is a way of characterising something that absorbs a lot of heat for a small change in its temperature).

Read about pseudo-explanations

Anthropomorphism explains by assuming that the entities involved can be considered to be like people, and, so, to be sentient, have feelings and opinions and preferences, and be able to plan and carry out actions that are intended to being about desired consequences.

It relies on an analogy that may not be appropriate:

  • if people were in a situation like this, they are likely to behave in a certain way
  • if we treat these entities as if they were people then we might expect them to behave as people would, therefore…

It is an open question to what extent we can assume animals (chimpanzees, dogs, birds, etc.) can be considered to share aspects of human-like experiences, emotions, thoughts, etcetera. Perhaps it is reasonable to suggest a dog can be sad or a chimp can be jealous. It may not be stretching credibility to suggest that members of some species of animals want to be in large groups, like to be in large groups, and perhaps may even get nervous when isolated? However, it stretches credibility when we are told that viruses are clever or that a bacterium can be happy.

And, there is a pretty strong scientific consensus that at the level of individual molecules there is no possibility of emotions, opinions, desires, thoughts, or deliberate actions. Atoms do not want to fill their electron shells, and genes cannot be selfish, except in a figurative sense.

Read about anthropomorphism

So, in order to accept Feynman's explanation as valid, we would have to assume that the quanticles in water, water molecules,

  • like to be next to each other
  • want as many partners as they can get 2
  • can be unhappy and nervous
  • try to have as many partners as possible 2
  • try to get into the inside of the volume

So, to find this explanation convincing, we have to accept (contrary to science) that something like a water molecule is able to

  1. have desires and preferences,
  2. be aware of the extent to which is current situation matches its preferences, and,
  3. deliberately act to bring about desired outcomes

[Feynman does not explicitly state that the quanticles know about their situation (point 2), but clearly this is implied as otherwise they would have no reason to be nervous and unhappy, nor to act to bring about change.]

These requirements are clearly not met. A being with a central nervous system as complex as a human can meet these requirements, but there is no conceivable mechanism by which molecules can be considered sentient, or to be deliberate agents in the world.

So, even if Feynman's explanation of surface tension satisfies viewers of the recording (i.e., is is subjectively an effective explanation) it fails as an objective, scientific, explanation. Feynman may indeed have been a 'genius' (Gleick, 1994), and a great physicist, but his explanation here is invalid and simply fails as good science.

Now a reader may suspect I have gone after a 'straw man' target here. Surely, Feynman was speaking figuratively, and not literally. Of course he was, but figurative language cannot support a logical explanation, except through an analogy we suspect to hold.

Consider the following hypothetical claim and two possible consequences if the claim was true

ClaimConsequence 1Consequence 2
"I managed to get tickets for Toyah and Fripp's sold out concert in Bury St Edmunds, and these tickets are gold dust.""I could sell these tickets at quite a mark up""I could put a sample of these tickets in a mass spectrometer and would find they had an atomic mass of 197."

If the claim was literally true, then consequence 2 would follow. But of course, it is meant as a figurative claim, where 'gold dust' is a metaphor for something of high value because it is rare. So, actually consequence 1 might follow, but not consequence 2.

In the same way, if water particles do not have likes, and do not try to do things, Fenyman's argument seems to fall apart…

A teaching model?

Now I would not presume to know better than Richard Feynman, and I am pretty sure (i.e., about as certain as I can be of anything) that Feynman would not have fallen into the mistake of thinking that atoms or molecules actually act like humans and want things, or try to do things. He surely knew this was not a scientific explanation, but he clearly thought this was a useful way of explaining (to his audience) why water forms into a drop.

Now, I suggested above that Feynman's narrative account of the origin of surface tension "is not based on accepted scientific principles (…or at least, not directly)". But near the outset of this account Feynman states that the water particles "attract each other":

"the [molecules] attract each other, they like to be next to each other"

Feynman was not only a researcher, but a teacher, and teachers use teaching models. I think this is what Feynman is doing here:

"[according to science] the [molecules] attract each other [and we can think of this as if] they like to be next to each other"

Affinity in the sense of human experience is used as a kind of analogy for the affinity between water molecules (which leads to hydrogen bonding and dipole-dipole interactions). Once we model inter-molecular forces as being like attractions between people, we can extend the analogy in terms of how people feel when they do not get what they want, and how they respond by acting in ways that they hope will get them what they want.

Looked at this way, Feynman is doing something that good teachers often do when they judge a scientific model is too abstract, sophisticated, complex, subtle, for their audience; they simplify by substituting a teaching model which represents the scientific model in terms more familiar and accessible to the learners.

Read about making the unfamiliar familiar

From this perspective, Feynman's explanation may not have been a valid scientific explanation, but we might ask if it was an effective intermediate explanation set out in terms of a teaching model. That is, perhaps Feynman's explanation may have satisfied viewers, and also potentially acted as a possible foundation for building up to a more technical, scientifically acceptable explanation.

Teachers and other science communicators often use anthropomorphism as a way of offering accounts of complex scientific topics that will appeal and make sense to learners of a public audience.

Read about anthropomorphism in accounts of science

This can be effective to the extent that it engages learners, leaves the audience with a subjective sense of making sense of the science, and provides accounts that are often remembered later.

Of course that is not so helpful if the audience is studying a science course, and think they have learnt an account which will get them credit in formal examinations! I know from my own teaching career that learners often find anthropomorphic explanations readily come to mind, even when then they have been taught more technical accounts they are expected to apply when assessed.

In public science communication, then, anthropomorphic accounts may be valuable in offering people some sense of the science. But in formal education we need to be careful as even if anthropomorphism offers a useful way of getting learners familiar with some abstract topic (what might be called 'weak' anthropomorphism: Taber & Watts, 1996), we need to avoid them learning and committing to that metaphoric 'social' account thinking it is a valid scientific account ('strong' anthropomorphism).

Mapping Feynman's explanation

If we see Feynman as offering an analogy as a teaching model then we might seek to 'translate' his terms into more scientific concepts. He tells us that attraction is 'liking', and we can perhaps think of 'wanting' and being 'nervous' as indicating a higher (excited) energy state, 'pounding' as being subject to unbalanced forces, and 'trying to get in' as tending to evolving toward a lower energy configuration. At least, someone who already understood the scientific account could suggest such mappings. It seems unlikely any one who did not appreciate the science already could interpret it that way without a knowing and careful guide.

And like all anthropomorphic explanations, it 'suffers' from the very quality that it offers a narrative which is likely to be more easily understood, better related to, and more readily recalled, than the scientific account. This is why I have very mixed feelings about the use of anthropomorphism in formal science teaching, as even when it (a) does a great job of engaging learners and offering them some level of understanding, this may be at the cost of (b) offering an account which many students will find it hard to later let go of and progress beyond.

Screenshot of Richard Feynman explaining why water forms into drops.


As a good teacher, Feynman would know to pitch his teaching for particular audiences depending on their likely level of background knowledge. The explanation discussed here was not how Feynman taught about surface tension in his undergraduate classes at the California Institute of Technology (Feynman, Leighton & Sands, 1963). We can imagine that had he told students at Caltech that water formed into spherical drops because all the molecular guys are trying to get into the water, he might indeed had heard the retort: Surely you are joking, Prof. Feynman? 1


Work cited:

Notes:

1 My subtitle is a reference to the book 'Surely you're Joking Mr Feynman: Adventures of a Curious Character' in which Feynman tells anecdotes from his life.


2 Water was perhaps a poor example to choose as there is extensive hydrogen bonding in liquid water,

"I suspect even some experienced chemists may underestimate the extent of hydrogen bonding in water. According to one source…, in liquid water at the freezing point, the typical water molecule is at any time bonded by three or four hydrogen bonds – compared with the four bonds in the solid ice structure."

Taber, 2020, p.98

So, in Feynman's analogy, a water molecules does not become happy (lower energy state) when it is surrounded by as many other water molecules as possible, but when it is aligned with 3 or 4 other molecules to hydrogen bond, if only transiently. Without the hydrogen bonding, the drop would still be approximately spherical, but it would be smaller and denser as the molecules could get even closer together, but it would evaporate away more readily.


Are the particles in all solids the same?

Particle intuitions may not match scientific models


Keith S. Taber


Sophia was a participant in the Understanding Science Project. I first talked to her when she was in Y7, soon after she began her secondary school course.

One of the first topics she studied in her science was 'solids, liquids and gases', where she had learnt,

that solids are really hard and they stay together more, and then liquids are close together but they move around, and gases are really free and they just go anywhere

She had studied a little about the topic in her last year of primary school (Y6), but now she was being told

about the particles…the things that make – the actual thing, make them a solid, and make them a gas and make them a liquid

Particle theory, or basic kinetic theory, is one of the most fundamental theories of modern science. In particular, much of what is taught in school chemistry is explained in terms of theories involving how the observed macroscopic properties emerge from the characteristics and interactions of conjectured sub-microscopic particles that themselves often have quite unfamiliar properties. This makes the subject very abstract, challenging, and tricky to teach (Taber, 2013a).

Read about conceptions of atoms

Particle theory is often introduced in terms of the states of matter. Strictly there are more than three states of matter (plasma and Bose-Einstein condensates are important in some areas of science) but the familiar ones, and the most important in everyday phenomena, are solid, liquid and gas.

The scientific account is, in simple terms, that

  • different substances are made up of different types of particle
  • the different states of matter of a single substance have the same particles arranged differently

These are very powerful ideas, even if there are many complications. For example,

  • the terms solid, liquid and gas only strictly apply to pure samples of a single substance, not mixtures (so not, for example, to bronze, or honey, or, milk, or ketchup, or even {if one is being very pedantic} air or sea water. And cats (please note, BBC) are completely inadmissible. )
  • common salt is an example of a pure substance, that none-the-less is considered to be made up of more than one type of particle

This reflects a common type of challenge in teaching science – the full scientific account is complex and nuanced, and not suitable for presenting in an introductory account; so we need to teach a simplified version that introduced the key ideas, and then only once this is mastered by learners are they ready to develop a more sophisticated understanding.

Yet, there is a danger that students will learn the simplified models as truths supported by the authority of science – and then later have difficulty shifting their thinking on. This is not only counter-productive, but can be frustrating and de-motivating for learners who find hard-earned knowledge is not as sound as they assumed.

One response to this is to teach science form very early in a way that is explicit about how science builds models of the natural world: models that are often simplifications which are useful but need to be refined and developed to become powerful enough to expand the range of contexts and examples where they can be applied. That is, students should learn they are being taught models that are often partial or imperfect, but that is just a reflection of how science works, developing more sophisticated understanding over time (Taber, 2017).

Sophia confirmed that the iron clamp stand near where she was sitting would have particles in it, as would a lump of ice.

Are they the same particles in the ice as the iron?

Yeah, because they are a solid, but they can change.

Ah, how can they change?

Cause if, erm, they melted they would be a liquid so they would have different particles in.

Right, so the iron is a solid, 

Uh hm.

So that's got one type of particle?

Yeah.

And ice is also a solid?

Yeah.

So that has the same sort of particles?

Yeah, but they can change.

The ones in the ice?

Mm,


To a learner just meeting particle theory for the first time, it may seem just as feasible that the same type of particle is found in one state as in one substance.


In the scientific model, we explain that different substances contain different types of particles, whereas different states of the same substance contain different arrangements of the same particles: but this may not be intuitively obvious to learners.1 It seemed Sophia was thinking that the same particles would be in different liquids, but a change of state led to different particles. This may seem a more forced model to a teacher, but then the teacher is already very familiar with the scientific account, and also has an understanding of the nature of those particles (molecules, ions, atoms – with internal structure and charges that interact with each other within and between the particles) – which are just vague, recently imagined, entities to the novice.

Sophia seemed to misunderstood or misremembered the model she had been taught, but to a novice learner these 'particles' have no more immediate referent than an elf or an ogre and would be considerably more tenuous than a will-o'-the-wisp.

Sophia seemed to have an alternative conception, that all solids have one type of particle, and all liquids another. If I had stopped probing at that point I might have considered this to be her thinking on the matter. However, when one spends time talking to students it soon becomes clear that often they have ideas that are not fully formed, or that may be hybrids of different models under consideration, and that often as they talk they can talk themselves into a position.

So, if I melted the ice – that changes the particles in the solid?

Well they are still the same particles but they are just changing the way they act…

Oh.

How do they change?

A particle in a liquid [sic, solid] is all crammed together and don't move around, but in a liquid they can move around a little but they are still close and, can, you can pour a liquid, where you can't a solid, because they can move in. 

Okay, so if I have got my ice, that's a solid, and there are particles in the ice, and they behave in a certain way, and if the ice melts, the particles behave differently?

Yeah.

Do you know why they behave differently in the liquid?

No. {giggles} So, they can, erm

• • • • • • • • • • • •  [A pause of approximately 12 s]

They've more room cause it's all spread out more1, whereas it would be in a clump

The literature on learners conceptions often suggests that students have this or that conception, or (when survey questions are used) that this percentage thinks this, and that percentage thinks that (Taber, 2013b). That this is likely to be a simplification seems obvious is we consider what thinking is – whatever thought may be, is it a dynamic process, something that moves along. Our thinking is, in part, resourced by accessing what we have represented in memory, but it is not something fixed – rather something that shifts, and that often becomes more sophisticated and nuanced as we explore a focus in greater depth.

I think Sophia did seem to have an intuition that there were different types of particles in different states of matter, and that therefore a change of state meant the particles themselves changed in some way. As I probed her, she seemed to shift to a more canonical account where change of state involved a change in the arrangement or organisation of particles rather than their identity.

This may have simply been her gradually bringing to mind what she had been taught – remembering what the teacher had said. It is also possible that the logic of the phenomenon of a solid becoming a liquid impressed on her that they must be the same particles. I suspect there was a little of both.

When interviewing students for research we inevitably change their thinking and understanding to some extent (hopefully, mostly in a beneficial way!) (If only teachers had time to engage each of their students in this way about each new topic they might both better understand their students' thinking, and help reinforce what has been taught.)

Did Sophia 'have a misconception'? 1 What did she 'really think'? That, surely, is to oversimplify.

She presented with an alternative conception, that under gentle questioning she seemed to talk /think herself out of. The extent to which her shift in position reflected further recall (so, correcting her response) or 'thinking through' (so, developing her understanding) cannot be known. Likely there was a little of both. What memory research does suggest is that being asked to engage in and think about this material will have modified and reinforced her memories of the material for the future.

Read about the role of memory in teaching and learning


Work cited:

Note

1 Actually, the particles in a liquid are not substantially spread further apart than in a solid. (Indeed, when ice melts the water molecules move closer together on average.) Understanding melting requires an appreciation of the attractions between particles, and how heating provides more energy for the particles. This idea of increased separation on melting is therefore something of an alternative conception, if one that is sometimes encouraged by the diagrams in school textbooks.

Teaching an introductory particle theory based on the arrangement of particles in different states, without reference to the attractions between particles is problematic as it offers no rational basis for why condensed states exists, and why energy is needed to disrupt them – something highlighted in the work of Philip Johnson (2012).



Scientific errors in the English National Curriculum

Keith S. Taber

I am writing this open letter to the Institute of Physics and the Royal Society of Chemistry to request that as Learned Societies with some influence with government (perhaps limited, but certainly vastly more than an academic) the Societies might ask the Department for Education to correct two basic errors of science in the National Curriculum for England which is set out as the basis for teaching school age learerns and for developing public examinations specifications and papers.
The two errors relate to (a) the misuse of scientific terminology (the word substance) and (b) a failure of logic (in a reference to conservation of energy). As you will no doubt be aware, the original published version of this iteration of the programmes of study for science in the English National Curriculum included some basic errors (incorrect physics formulae) that received wide publicity and which were quickly amended. Despite some other issues also getting early attention, these other problems have never been addressed. One more complex issue that I strongly feel deserves addressing, but which would would require considerable redrafting, is the confused and incoherent treatment of the nature of chemical reactions across the secondary phase (Key Stages 3 and 4). I have raised these issues at various times, and have published a scholarly analysis of these problems .Whilst I obviously did not expect an article in an academic journal to directly impact policy, I thought this could be a 'springboard' to then approach government. I have contacted the relevant ministers (the Rt Hon Gavin Williamson CBE MP, Secretary of State for Education and the Rt Hon Nick Gibb MP, Minister of State for School Standards), and in response to instructions to refer this issue to the Department for Education website, I did so. My comments have been noted, but I was informed
"there are no current plans to review the curriculum".
Whilst I accept that any detailed re-working of the curriculum is not imminent, I do think the Department could still instigate minor corrections to errors which are published on the government's website, and then consequently repeated by the examination authorities, the examination boards and even individual school websites. Correcting these (surely, embarrassing) errors would require very little effort. The first error I refer to is the incorrect use of the term 'substance'. In science, the term substance has a fairly specific meaning. Although, as with many science concepts, there may be some discussion over precise definitions and demarcations, there is general agreement at the level at which the term would be used in introductory science at school level. In the primary stages of the English National Curriculum for Science we read that Y5 learners should be
"taught to…explain that some changes result in the formation of new materials [sic], and that this kind of change is not usually reversible, including changes associated with burning and the action of acid on bicarbonate of soda".
A better term here would be 'substances', not 'materials' (although this is more a mater of the wording being imprecise than incorrect). However in relation to Y4 learners there is a reference to
"exploring the effect of temperature on substances [sic] such as chocolate, butter, cream"
none of which are substances as the word is used in science.This is a misuse of the term 'substance'. So whereas in secondary school, learners are taught to distinguish the meanings of 'material' and the more specific 'substance', it seems these terms are being used interchangeably in the National Curriculum specification itself. The other issue relates to the statement (in the Key Stage 4 specification) that
"energy is conserved in chemical reactions so can therefore be neither created nor destroyed".
To my reading this suggests a blatant error of logic, which I can only assume does not reflect scientific ignorance by the person drafting the document – but more likely is a typographic error that has never been corrected. Conservation of energy is a general (universal) principle, and its more specific application to chemical reactions as one class of changes is then subsumed under that principle. I have long assumed that what had been intended (but mistyped) was either "energy is conserved in chemical reactions BECAUSE it can be neither created nor destroyed" or "energy CAN be neither created nor destroyed SO THEREFORE is conserved in chemical reactions" – that is, the logic has been completely reversed in the curriculum document. I have recently realised that there is a third possibility: that this statement is not meant as an explanation (of energy conservation in reactions under a more general principle) but as a definition, along the lines "energy is conserved in chemical reactions WHICH MEANS THAT IT CAN be neither created nor destroyed". Whatever was meant, the current wording implies a logical non sequitur, and should, surely, be corrected. I would hope you might agree that these kinds of errors should not be included in what teachers are asked to teach, students to learn, and examining boards to assess; and that when a suitable opportunity arrises you might make appropriate representations regarding the desirability of corrections being made. Your sincerely, Dr Keith S.Taber Emeritus Professor of Science Education (I have had constructive replies from both the RSC and IoP)

Thank you, BBC: I'll give you 4/5

BBC corrects cruel (to cats) scientific claim on its website

Keith S. Taber

I just got 80% on a science test for primary school children

I've just scored 4/5 (80%) on an on-line KS2 science test on the BBC (the British Broadcasting Corporation) educational website. 80% sounds quite good out of context, but I am a science teacher and KS2 is meant for 7-11 year olds.

The BBC awards me 4/5 for my primary level science knowledge about the states of matter

My defence is that the question I got wrong was ambiguous (but, as Christine Keeler might have said, I would say that).

I was actually getting round to checking on something from a while back.

In 2019 I came across something on the website that I thought was very misleading – and I complained to the BBC through their website form. I had an immediate, but generic response:

"Thank you for taking the time to send us your comments. We appreciate all the feedback we receive as it plays an important role in helping to shape our decisions.

This is an automated message (sorry that we can't reply individually) to let you know that we've read your comments and will report them overnight to staff across the BBC for them to read too (after removing any personal details). This includes our programme makers, commissioning editors and senior management.

Thanks again for contacting the BBC.

BBC Audience Services.

NB: Please do not reply to this email. It includes a reference number but comes from an automated account which is not monitored."

Email: 6th Sept., 2019

This kind of response is somewhat frustating. My complaint had been recieved, and would be passed on, but it looked like I would get no specific response (as presumably if my "comments" were to be reported to relevant staff "after removing any personal details", those staff would not be in a position to let me know if they were following up, dismissing, or simply ignoring, my comments.) Indeed, I never did get any follow up.

So, my intention was to check back after a decent period had elapsed (n.b., where does all the time go?) and see if anything had been changed in response to my complaint. Strictly, if there had been a change this could be because:

  • a) I complained
  • b) someone else/some other people complained (i.e., people who's complaints were taken more seriously than mine)
  • c) I was one of number of people who complained
  • d) material had been updated compleltely independently of any compaints

That is, I could not know if I personally had had any effect, BUT if the offending material (because as a chemist I was offended professionally, even if not personally) was still there then I would know my compaint had not been heeded.

So, I intended to check back; I expected to find no change (as pointing out blatant, basic, errors in the science in the English National Curriculum to government ministers did not have any effect, so the BBC…? ); and, if so, I thought of following up with an email or an old fashioned snail-mail … ("…yours, disgusted of Cambourne"*).

Well done, BBC

So, I am happy to publicly acknowledge that the BBC has changed its materials appearing under the heading 'What are the states of matter?'

The topic comprises of a short animation (with odd anthropomorphised {"guys"} geometric shapes handling examples of the states of matter: solid, liquid and gas); a series of bullet points on each state; a sorting task; and then the set of five objective (multiple choice) questions.

There are a number of issues with the examples used here, as discussed below. But the main focus of my complaint, a cartoon cat, has now been released from the indignity of being classified as a state of matter. Yes, a cat!

Limitations of the three states of matter model

The idea that matter can exist in three states is a pretty important foundation for a good deal of other science.

However there is big problem with the generality of the model. Basically it really applies to pure samples of substances: generally substances (not materials in general, and certainly not objects) exist as solids, liquids, or gases, depending on the conditions of temperature and pressure – although at high enough temperatures plasmas are formed (and theoretically when hot enough even the atomic cores, and eventually nuclei would break down – but those conditions are pretty extreme and not found in the typical home or classroom).

Examples of substances include water, salt, calcium carbonate, iron, mercury, hydrogen, graphite, carbon dioxide, sulphur… that is, elements and compounds. Of course, many of these are seldom met in pure form in everyday life outside school science labs.

Most materials that people come across are mixtures or composites. Mixtures often exist as solutions or suspensions – as gels or foams or emulsions – not as solids, liquids or gases.

This is probably why the terms 'solids', 'liquids' and 'gases' actually have two sets of meanings – the science or technical sense, and the everyday or 'life-world' sense. So milk is a liquid(everyday) as you can pour some into your tea cup and a block of wood is a solid(everyday) as it retains its shape and integrity as you nail it to another structure. But milk and wood are not substances – and so not liquid(scientific) or solid(scientific).

Does this matter? Yes, because if we are teaching children things in science lessons, it would be good to get the science right. A solid will melt at a distinct melting temperature to give a liquid which will boil at a distinct boiling temperature. Wood, for example, does not.

Wood is a complex material. It has gas pockets. It has (variable) moisture content, and the structure contains various compounds – lignin, cellulose, and many more. The response to heating reflects that complex constitution.

The BBC's examples of solids, liquids, and gases

The BBC website suggests examples of the three states of matter to introduce primary age students to the concept.

Animation:

Solids: block of ice, football

Lquids: water, honey

Gases: none are specified – animation shows the clouds (of liquid water droplets) forming around a kettle spout, and 'gas' put into in fizzy drinks is referenced.

A football is not solid, but usually air (a mixture of gases with some other components) contained in a plastic shell. (The voiceover refers simply to a 'ball', but the animation show a large ball with a traditional football pattern being used to do 'keepy uppies' by the cartoon character.)

Honey is not a liquid(scientific) but a complex mixture of sugars in solution. There is usually much more sugar than water. (So, arguably, it is more solid than liquid – but it is better to simply not consider it as either.) This is where I dropped a mark on the terminal test:

Two of the options are NOT liquids. Only one response gets credit in this test!
Web text:

The bullet points on the site list some further examples:

"Examples of solids include ice, wood and sand." (Ice and sand are solids(scientific).)

"Examples of liquids include water, honey and milk." (Only water is liquid(scientific) here.)

"Examples of gases include steam, helium and oxygen." (3/3, well done BBC!)

Sorting task:
The BBC website task invites children to sort cards showing objects into three categories. (What is that object on the front card meant to be?)

In the sorting task, children are asked to sort a number of examples shown on cards into solid, liquid, and gas:

The examples presented are air, a feather, helium, milk, a pencil, sea, steam, syrup, wood. Of these only helium and steam strictly meet the criteria for being a solid(scientific)/liquid(scientific)/gas(scientific). Yet, as suggested above, it is difficult to find genuine examples that are both scientifically correct and familiar to young children. Perhaps sea and air (at least materials) are closer approximations than a pencil or a feather ("solids retain their shape" – would a child using the website have handled a feather, and, if so, would it have retained its shape under child-handling?)

So, I still have reservations about this material, whilst acknowledging the need to balance scientific correctness with relevant (to children) examples. Strictly, some of the examples can be seen as encouraging children to get the science wrong. These things matter if only because children are learning things on this site that later in their school career will be judged as alternative conceptions and marked as wrong.

(Read 'Are plants solid?')

None the less, I am pleased that the BBC has at least decided to amend its sorting task, and remove the poor cat:

Which pile does the cat belong in? [This example has now been removed. Bravo.]

The website had previously been quite clear that putting the cat as anything other than solid was 'wrong'. It is classed as a solid even though a cat (like any animal) is (or would be if separated out into its constituent substances – and children should not try this at home) more water than anything else.

I had real trouble seeing how that example fitted with the criteria specified on the webpage:

"[Cats] stay in one place and can be held.

[Cats] keep their shape. They do not flow like liquids.

[Cats] always take up the same amount of space. They do not spread out like gases.

[Cats] can be cut or shaped."

Characteristics of solids, but perhaps not entirely true of cats?

* cf. the idiom 'disgusted of Tunbridge Wells' – referring to a hypothetical person who writes to media complaining about matters of concern.

Images used here are screenshots, copyright of the BBC – a publicly funded public service broadcaster.

A case of hybrid research design?

When is "a case study" not a case study? Perhaps when it is (nearly) an experiment?

Keith S. Taber

I read this interesting study exploring learners shifting conceptions of the particulate nature of gases.

Mamombe, C., Mathabathe, K. C., & Gaigher, E. (2020). The influence of an inquiry-based approach on grade four learners' understanding of the particulate nature of matter in the gaseous phase: a case study. EURASIA Journal of Mathematics, Science and Technology Education, 16(1), 1-11. doi:10.29333/ejmste/110391

Key features:

  • Science curriculum context: the particulate nature of matter in the gaseous phase
  • Educational context: Grade 4 students in South Africa
  • Pedagogic context: Teacher-initiated inquiry approach (compared to a 'lecture' condition/treatment)
  • Methodology: "qualitative pre-test/post-test case study design" – or possibly a quasi-experiment?
  • Population/sample: the sample comprised 116 students from four grade four classes, two from each of two schools

This study offers some interesting data, providing evidence of how students represent their conceptions of the particulate nature of gases. What most intrigued me about the study was its research design, which seemed to reflect an unusual hybrid of quite distinct methodologies.

In this post I look at whether the study is indeed a case study as the authors suggest, or perhaps a kind of experiment. I also make some comments about the teaching model of the states of matter presented to the learners, and raise the question of whether the comparison condition (lecturing 8-9 year old children about an abstract scientific model) is appropriate, and indeed ethical.

Learners' conceptions of the particulate nature of matter

This paper is well worth reading for anyone who is not familiar with existing research (such as that cited in the paper) describing how children make sense of the particulate nature of matter, something that many find counter-intuitive. As a taster for this, I reproduce here two figures from the paper (which is published open access under a creative common license* that allows sharing and adaption of copyright material with due acknowledgement).

Figures © 2020 by the authors of the cited paper *

Conceptions are internal, and only directly available to the epistemic subject, the person holding the conception. (Indeed, some conceptions may be considered implicit, and so not even available to direct introspection.) In research, participants are asked to represent their understandings in the external 'public space' – often in talk, here by drawing (Taber, 2013). The drawings have to be interpreted by the researchers (during data analysis). In this study the researchers also collected data from group work during learning (in the enquiry condition) and by interviewing students.

What kind of research design is this?

Mamombe and colleagues describe their study as "a qualitative pre-test/post-test case study design with qualitative content analysis to provide more insight into learners' ideas of matter in the gaseous phase" (p. 3), yet it has many features of an experimental study.

The study was

"conducted to explore the influence of inquiry-based education in eliciting learners' understanding of the particulate nature of matter in the gaseous phase"

p.1

The experiment compared two pedagogical treatments :

  • "inquiry-based teaching…teacher-guided inquiry method" (p.3) guided by "inquiry-based instruction as conceptualized in the 5Es instructional model" (p.5)
  • "direct instruction…the lecture method" (p.3)

These pedagogic approaches were described:

"In the inquiry lessons learners were given a lot of materials and equipment to work with in various activities to determine answers to the questions about matter in the gaseous phase. The learners in the inquiry lessons made use of their observations and made their own representations of air in different contexts."

"the teacher gave probing questions to learners who worked in groups and constructed different models of their conceptions of matter in the gaseous phase. The learners engaged in discussion and asked the teacher many questions during their group activities. Each group of learners reported their understanding of matter in the gaseous phase to the class"

p.5, p.1

"In the lecture lessons learners did not do any activities. They were taught in a lecturing style and given all the notes and all the necessary drawings.

In the lecture classes the learners were exposed to lecture method which constituted mainly of the teacher telling the learners all they needed to know about the topic PNM [particulate nature of matter]. …During the lecture classes the learners wrote a lot of notes and copied a lot of drawings. Learners were instructed to paste some of the drawings in their books."

pp.5-6

The authors report that,

"The learners were given clear and neat drawings which represent particles in the gaseous, liquid and solid states…The following drawing was copied by learners from the chalkboard."

p.6
Figure used to teach learners in the 'lecture' condition. Figure © 2020 by the authors of the cited paper *
A teaching model of the states of matter

This figure shows increasing separation between particles moving from solid to liquid to gas. It is not a canonical figure, in that the spacing in a liquid is not substantially greater than in a solid (indeed, in ice floating on water the spacing is greater in the solid), whereas the difference in spacing in the two fluid states is under-represented.

Such figures do not show the very important dynamic aspect: that in a solid particles can usually only oscillate around a fixed position (a very low rate of diffusion not withstanding), where in a liquid particles can move around, but movement is restricted by the close arrangement of (and intermolecular forces between) the particles, where in a gas there is a significant mean free path between collisions where particles move with virtually constant velocity. A static figure like this, then, does not show the critical differences in particle interactions which are core to the basic scientific model

Perhaps even more significant, figure 2 suggests there is the same level of order in the three states, whereas the difference in ordering between a solid and liquid is much more significant than any change in particle spacing.

In teaching, choices have to be made about how to represent science (through teaching models) to learners who are usually not ready to take on board the full details and complexity of scientific knowledge. Here, Figure 2 represents a teaching model where it has been decided to emphasise one aspect of the scientific model (particle spacing) by distorting the canonical model, and to neglect other key features of the basic scientific account (particle movement and arrangement).

External teachers taught the classes

The teaching was undertaken by two university lecturers

"Two experienced teachers who are university lecturers and well experienced in teacher education taught the two classes during the intervention. Each experienced teacher taught using the lecture method in one school and using the teacher-guided inquiry method in the other school."

p.3

So, in each school there was one class taught by each approach (enquiry/lecture) by a different visiting teacher, and the teachers 'swapped' the teaching approaches between schools (a sensible measure to balance possible differences between the skills/styles of the two teachers).

The research design included a class in each treatment in each of two schools

An experiment; or a case study?

Although the study compared progression in learning across two teaching treatments using an analysis of learner diagrams, the study also included interviews, as well as learners' "notes during class activities" (which one would expect would be fairly uniform within each class in the 'lecture' treatment).

The outcome

The authors do not consider their study to be an experiment, despite setting up two conditions for teaching, and comparing outcomes between the two conditions, and drawing conclusions accordingly:

"The results of the inquiry classes of the current study revealed a considerable improvement in the learners' drawings…The results of the lecture group were however, contrary to those of the inquiry group. Most learners in the lecture group showed continuous model in their post-intervention results just as they did before the intervention…only a slight improvement was observed in the drawings of the lecture group as compared to their pre-intervention results"

pp.8-9

These statements can be read in two ways – either

  • a description of events (it just happened that with these particular classes the researchers found better outcomes in the enquiry condition), or
  • as the basis for a generalised inference.

An experiment would be designed to test a hypothesis (this study does not seem to have an explicit hypothesis, nor explicit research questions). Participants would be assigned randomly to conditions (Taber, 2019), or, at least, classes would be randomly assigned (although then strictly each class should be considered as a single unit of analysis offering much less basis for statistical comparisons). No information is given in the paper on how it was decided which classes would be taught by which treatment.

Representativeness

A study could be carried out with the participation of a complete population of interest (e.g., all of the science teachers in one secondary school), but more commonly a sample is selected from a population of interest. In a true experiment, the sample has to be selected randomly from the population (Taber, 2019) which is seldom possible in educational studies.

The study investigated a sample of 'grade four learners'

In Mamombe and colleagues' study the sample is described. However, there is no explicit reference to the population from which the sample is drawn. Yet the use of the term 'sample' (rather than just, say, 'participants') implies that they did have a population in mind.

The aim of the study is given as to "to explore the influence of inquiry-based education in eliciting learners' understanding of the particulate nature of matter in the gaseous phase" (p.1) which could be considered to imply that the population is 'learners'. The title of the paper could be taken to suggest the population of interests is more specific: "grade four learners". However, the authors make no attempt to argue that their sample is representative of any particular population, and therefore have no basis for statistical generalisation beyond the sample (whether to learners, or to grade four learners, or to grade four learners in RSA, or to grade four learners in farm schools in RSA, or…).

Indeed only descriptive statistics are presented: there is no attempt to use tests of statistical significance to infer whether the difference in outcomes between conditions found in the sample would probably have also been found in the wider population.

(That is inferential stats. are commonly used to suggest 'we found a statistically significant better outcome in one condition in our sample, so in the hypothetical situation that we had been able to include the entire population in out study we would probably have found better mean outcomes in that same condition'.)

This may be one reason why Mamombe and colleagues do not consider their study to be an experiment. The authors acknowledge limitations in their study (as there always are in any study) including that "the sample was limited to two schools and two science education specialists as instructors; the results should therefore not be generalized" (p.9).

Yet, of course, if the results cannot be generalised beyond these four classes in two schools, this undermines the usefulness of the study (and the grounds for the recommendations the authors make for teaching based on their findings in the specific research contexts).

If considered as an experiment, the study suffers from other inherent limitations (Taber, 2019). There were likely novelty effects, and even though there was no explicit hypothesis, it is clear that the authors expected enquiry to be a productive approach, so expectancy effects may have been operating.

Analytical framework

In an experiment is it important to have an objective means to measure outcomes, and this should be determined before data are collected. (Read about 'Analysis' in research studies.). In this study methods used in previous published work were adopted, and the authors tell us that "A coding scheme was developed based on the findings of previous research…and used during the coding process in the current research" (p.6).

But they then go on to report,

"Learners' drawings during the pre-test and post-test, their notes during class activities and their responses during interviews were all analysed using the coding scheme developed. This study used a combination of deductive and inductive content analysis where new conceptions were allowed to emerge from the data in addition to the ones previously identified in the literature"

p.6

An emerging analytical frame is perfectly appropriate in 'discovery' research where a pre-determined conceptualisation of how data is to be understood is not employed. However in 'confirmatory' research, testing a specific idea, the analysis is operationalised prior to collecting data. The use of qualitative data does not exclude a hypothesis-testing, confirmatory study, as qualitative data can be analysed quantitatively (as is done in this study), but using codes that link back to a hypothesis being tested, rather than emergent codes. (Read about 'Approaches to qualitative data analysis'.)

Much of Mamombe and colleagues' description of their work aligns with an exploratory discovery approach to enquiry, yet the gist of the study is to compare student representations in relation to a model of correct/acceptable or alternative conceptions to test the relative effectiveness of two pedagogic treatments (i.e., an experiment). That is a 'nomothetic' approach that assumed standard categories of response.

Overall, the author's account of how they collected and analysed data seem to suggest a hybrid approach, with elements of both a confirmatory approach (suitable for an experiment) and elements of a discovery approach (more suitable for case study). It might seem this is a kind of mixed methods study with both confirmatory/nomothetic and discovery/idiographic aspects – responding to two different types of research question the same study.

Yet there do not actually seem (**) to be two complementary strands to the research (one exploring the richness of student's ideas, the other comparing variables – i.e., type of teaching versus degree of learning), but rather an attempt to hybridise distinct approaches based on incongruent fundamental (paradigmatic) assumptions about research. (** Having explicit research questions stated in the paper could have clarified this issue for a reader.)

So, do we have a case study?

Mamombe and colleagues may have chosen to frame their study as a kind of case study because of the issues raised above in regard to considering it an experiment. However, it is hard to see how it qualifies as case study (even if the editor and peer reviewers of the EURASIA Journal of Mathematics, Science and Technology Education presumably felt this description was appropriate).

Mamombe and colleagues do use multiple data sources, which is a common feature of case study. However, in other ways the study does not meet the usual criteria for case study. (Read more about 'Case study'.)

For one thing, case study is naturalistic. The method is used to study a complex phenomena (e.g., a teacher teaching a class) that is embedded in a wider context (e.g., a particular school, timetable, cultural context, etc.) such that it cannot be excised for clinical examination (e.g., moving the lesson to a university campus for easy observation) without changing it. Here, there was an intervention, imposed from the outside, with external agents acting as the class teachers.

Even more fundamentally – what is the 'case'?

A case has to have a recognisable ('natural') boundary, albeit one that has some permeability in relation to its context. A classroom, class, year group, teacher, school, school district, etcetera, can be the subject of a case study. Two different classes in one school, combined with two other classes from another school, does not seem to make a bounded case.

In case study, the case has to be defined (not so in this study); and it should be clear it is a naturally occurring unit (not so here); and the case report should provide 'thick description' (not provided here) of the case in its context. Mamombe and colleagues' study is simply not a case study as usually understood: not a "qualitative pre-test/post-test case study design" or any other kind of case study.

That kind of mislabelling does not in itself does not invalidate research – but may indicate some confusion in the basic paradigmatic underpinnings of a study. That seems to be the case [sic] here, as suggested above.

Suitability of the comparison condition: lecturing

A final issue of note about the methodology in this study is the nature of one of the two conditions used as a pedagogic treatment. In a true experiment, this condition (against which the enquiry condition was contrasted) would be referred to as the control condition. In a quasi-experiment (where randomisation of participants to conditions is not carried out) this would usually be referred to as the comparison condition.

At one point Mamombe and colleagues refer to this pedagogic treatment as 'direct instruction' (p.3), although this term has become ambiguous as it has been shown to mean quite different things to different authors. This is also referred to in the paper as the lecture condition.

Is the comparison condition ethical?

Parental consent was given for students contributing data for analysis in the study, but parents would likely trust the professional judgement of the researchers to ensure their children were taught appropriately. Readers are informed that "the learners whose parents had not given consent also participated in all the activities together with the rest of the class" (p.3) so it seems some children in the lecture treatment were subject to the inferior teaching approach despite this lack of consent, as they were studying "a prescribed topic in the syllabus of the learners" (p.3).

I have been very critical of a certain kind of 'rhetorical' research (Taber, 2019) report which

  • begins by extolling the virtues of some kind of active / learner-centred / progressive / constructivist pedagogy; explaining why it would be expected to provide effective teaching; and citing numerous studies that show its proven superiority across diverse teaching contexts;
  • then compares this with passive modes of learning, based on the teacher talking and giving students notes to copy, which is often characterised as 'traditional' but is said to be ineffective in supporting student learning;
  • then describes how authors set up an experiment to test the (superior) pedagogy in some specific context, using as a comparison condition the very passive learning approach they have already criticised as being ineffective as supporting learning.

My argument is that such research is unethical

  • It is not genuine science as the researchers are not testing a genuine hypothesis, but rather looking to demonstrate something they are already convinced of (which does not mean they could not be wrong, but in research we are trying to develop new knowledge).
  • It is not a proper test of the effectiveness of the progressive pedagogy as it is being compared against a teaching approach the authors have already established is sub-standard.

Most critically, young people are subjected to teaching that the researchers already believe they know will disadvantage them, just for the sake of their 'research', to generate data for reporting in a research journal. Sadly, such rhetorical studies are still often accepted for publication despite their methodological weaknesses and ethical flaws.

I am not suggesting that Mamombe, Mathabathe and Gaigher have carried out such a rhetorical study (i.e., one that poses a pseudo-question where from the outset only one outcome is considered feasible). They do not make strong criticisms of the lecturing approach, and even note that it produces some learning in their study:

"Similar to the inquiry group, the drawings of the learners were also clearer and easier to classify after teaching"

"although the inquiry method was more effective than the lecture method in eliciting improved particulate conception and reducing continuous conception, there was also improvement in the lecture group"

p.9, p.10

I have no experience of the South African education context, so I do not know what is typical pedagogy in primary schools there, nor the range of teaching approaches that grade 4 students there might normally experience (in the absence of external interventions such as reported in this study).

It is for the "two experienced teachers who are university lecturers and well experienced in teacher education" (p.3) to have judged whether a lecture approach based on teacher telling, children making notes and copying drawings, but with no student activities, can be considered an effective way of teaching 8-9 year old children a highly counter-intuitive, abstract, science topic. If they consider this good teaching practice (i.e., if it is the kind of approach they would recommend in their teacher education roles) then it is quite reasonable for them to have employed this comparison condition.

However, if these experienced teachers and teacher educators, and the researchers designing the study, considered that this was poor pedagogy, then there is a real question for them to address as to why they thought it was appropriate to implement it, rather than compare the enquiry condition with an alternative teaching approach that they would have expected to be effective.

Sources cited:

* Material reproduced from Mamombe, Mathabathe & Gaigher, 2020 is © 2020 licensee Modestum Ltd., UK. That article is an open access article distributed under the terms and conditions of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0/) [This post, excepting that material, is © 2020, Keith S. Taber.]

An introduction to research in education:

Taber, K. S. (2013). Classroom-based Research and Evidence-based Practice: An introduction (2nd ed.). London: Sage.

When is V=IR the formula for Ohm's law?

"Resistance is current over voltage, I think"

Image by Gerd Altmann from Pixabay 

Adrian was a participant in the Understanding Science Project. When I interviewed him in Y12 when he was studying Advanced level physics he told me that "We have looked at resistance and conductance and the formulas that go with them". So I asked him was resistance was:

So what exactly is resistance?

Resistance is, erm (pause, c.3s) Resistance is current over, voltage, I think. (Pause, c.3s) Yeah. No.

Not sure?

I can’t remember formulas.

So Adrian's first impulse was to define resistance using a formula, although he did not feel he was able to remember formulae. He correctly knew that the formula involved resistance, current and voltage, but could not recall the relationship. Of course if he understood qualitatively how these influenced each other, then he should have been able to work out which way the formula had to go, as the formula represents the relationship between resistance, voltage and current.

So, I then proceeded to ask Adrian how he would explain resistance to a younger person, and he suggested that resistance is how much something is being slowed down or is stopped going round. After we had talked about that for a while, I brought the discussion back to the formula and the relationship between R, V and I:

And what about this resistance in electricity then, do you measure that in some kind of unit?

Yes, in, erm, (pause, c.2s) In ohms.

So what is an ohm?

Erm, an ohm is, the unit that resistance is measured in.

Fair enough.

It comes from ohm's law which is the, erm, formula that gives you resistance.

V=IR is the formula that gives you resistance, but it is a common misconception, that Ohm's law is V=IR.

Actually, Ohm's law suggests that the current through a metallic conductor (kept at constant conditions, e.g., temperature) is directly proportional to the potential difference across its ends.

So, in such a case (a metal that is not changing temperature, etc.)

I ∝ V

which is equivalent to

V ∝ I

which is equivalent to

V = kI

where k is a constant of proportionality. If we use the symbol R for the constant in this case then

V= RI

which is equivalent to

V = IR

 So, it may seem I have just contradicted myself, as I denied that V=IR was a representation of Ohm's law, yet seem to have derived V=IR from the law.

There is no contradiction as long as we keep in mind what the symbols are representing in the equation. I represented the current flowing through a metallic conductor being held at constant conditions (temperature, tension etc.), and V represented the potential difference across the ends of that metallic conductor. If we restrict V and I to this meaning then the formula could be seen as a way of representing Ohm's law.

Over-generalising

However, that is not how we usually understand these symbols in electrical work: V generally represents the potential difference across some resistive component or other, and I represents the current flowing through that component: a resistor, a graphite rod, a salt bridge, a diode, a tungsten filament in a lamp…

In this general case

V = IR

or

R = V/I

is the defining equation for resistance. If R is defined as V/I then it will always be the case, not because there is a physical law that suggests this, but simply because that is the meaning we have given to R.

This is a bit like bachelors being unmarried men (an example that seems to be a favourite of some philosophers): bachelors are not unmarried men because there is some rule or law decreeing that bachelors are not able to get married, but simply because of our definition. A bachelor who gets married and so becomes a married man ceases to be a bachelor at the moment they become a married man – in a similar way to how a butterfly is no longer a caterpillar. Not because of some law of nature, but by our conventions regarding how words are used. If V and I are going to be used as general symbols (and not restricted to our carefully controlled metallic conductor) then V = IR simply because R is defined as V/I and the formula, used in the general case, should not be confused with Ohm's law.

In Ohm's law, V=IR where R will be constant.

In general, V=IR and R will vary, as Ohm's law does not generally apply.

It would perhaps be better for helping students see this had there been a convention that the p.d. across, and the current through, a piece of metal being kept in constant conditions were represented by, say V and I, so Ohm's law could be represented as, say

V = k I

but, as this is not the usual convention, students need to keep in mind when they are dealing with the special case to which Ohm's law refers.

A flawed teaching model?

The interesting question is whether:

  • teachers are being very careful to make this distinction, but students still get confused;
  • teachers are using language carefully, but not making the discrimination explicit for students, so they miss the distinction;
  • some teachers are actually teaching that V=IR is Ohm's law.

If the latter option is the case , then it would be good to know if the teachers teaching this:

  • have the alternative conception themselves;
  • appreciate the distinction, but think it does not matter;
  • consider identifying the general formula V=IR with Ohm's law is a suitable simplification, a kind of teaching model, suitable for students who are not ready to have the distinction explained.

It would be useful to know the answers to these questions, not to blame teachers, but because we need to diagnose a problem to suggest the best cure.



A wooden table is solid…or is it?

Keith S. Taber

Wood (cork coaster captured with Veho Discovery USB microscope)

Bill was a participant in the Understanding Science Project. Bill (Y7) was explaining that he had been learning about the states of matter, and introduced the notion of there being particles:

So how do you know if something is a solid, a liquid or a gas?

Well, solids they stay same shape and their particles only move a tiny bit

So what are these particles then?

Erm, they're the bits that make it what it is, I think.

Ah. So are there any solids round here?:

Yeah, this table. [The wooden table Bill was sitting at.]

That's a solid, is it?:

Yeah

Technically the terms solid, liquid and gas refer to samples of substances and not objects. From a chemical perspective a table is not solid. A wooden table (such as those in the school laboratory where I talked to Bill) is made of a complex composite material that includes various different substances such as a lignin and cellulose in its structure.

Wood contains some water, and has air pockets, so technically wood is not a solid to a chemist. However, in everyday life we do thing of objects such as tables as being solid.

Yet if wood is heated, water can be driven off. Timber can be mostly water by weight, and is 'seasoned' to remove much of the water content before being used as a construction material. Under the microscope the complex structure of woods can be seen, including spaces containing air.

Bill also suggested that a living plant should be considered a solid.

I think teaching may be a problem here, as when the states of matter are taught it is often not made clear these distinctions only apply clearly to fairly pure samples of substances. In effect the teaching model is that materials occur as solids, liquids and gases – when a good many materials (emulsions, gels, aerosols, etc.) do not fit this model at all well.

Particles are further apart in water than ice

Keith S. Taber

Image from Pixabay 

Bill was a participant in the Understanding Science Project. Bill, a Y7 student, explained what he had learnt about particles in solids, liquids and gases. Bill introduced the idea of particles when talking about what he had learn about the states of matter.

Well there's three groups, solids, liquids and gases.

So how do you know if something is a solid, a liquid or a gas?

Well, solids they stay same shape and their particles only move a tiny bit.

This point was followed up later in the interview.

So, you said that solids contain particles,

Yeah.

They don't move very much?

No.

And you've told me that ice is a solid?

Yeah.

So if I put those two things together, that tells me that ice should contain particles?

Yeah.

Yeah, and you said that liquids contain particles? Did you say they move, what did you say about the particles in liquids?

Er, they're quite, they're further apart, than the ones in erm solids, so they erm, they try and take the shape, they move away, but the volume of the water doesn't change. It just moves.

Bill reports that the particles in liquids are "further apart, than the ones in … solids". This is generally true* when comparing the same substance, but this is something that tends to be exaggerated in the basic teaching model often used in school science. Often figures in basic school texts show much greater spacing between the particles in a liquid than in the solid phase of the same material. This misrepresents the modest difference generally found in practice – as can be appreciated from the observations that volume increases on melting are usually modest.

Although generally the particles in a liquid are considered further apart than in the corresponding solid*, this need not always be so.

Indeed it is not so for water – so ice floats in cold water. The partial disruption of the hydrogen bonds in the solid that occurs on melting allows water molecules to be, on average, closer* in the liquid phase at 0˚C.

As ice float in water, it must have a lower density. If the density of some water is higher than that of the ice from which it was produced on melting then (as the mass will not have changed) the volume must have decreased. As the number of water molecules has not changed, then the smaller volume means the particles are on average taking up less space: that is, they seem to be closer together in the water, not further apart*.

Bill was no doubt aware that ice floats in water, but if Bill appreciated the relationship of mass and volume (i.e., density) and how relative density determines whether floatation occurs, he does not seem to have related this to his account here.

That is, he may have had the necessary elements of knowledge to appreciate that as ice floats, the particles in ice were not closer together than they were in water, but had not coordinated these discrete components to from an integrated conceptual framework.

Perhaps this is not surprising when we consider what the explanation would involve:

Coordinating concepts to understand the implication of ice floating

Not only do a range of ideas have to be coordinated, but these involve directly observable phenomena (floating), and abstract concepts (such as density), and conjectured nonobservable submicroscopic/nanoscopic level entities.

* A difficulty for teachers is that the entities being discussed as 'particles', often molecules, are not like familiar particles that have a definitive volume, and a clear surface. Rather these 'particles' (or quanticles) are fuzzy blobs of fields where the field intensity drops off gradually, and there is no surface as such.

Therefore, these quantiles do not actually have definite volumes, in the way a marble or snooker ball has a clear surface and a definite volume. These fields interact with the fields of other quanticles around them (that is, they form 'bonds' – such as dipole-dipole interactions), so in condensed phases (solids and liquids) there are really not any discrete particles with gaps between them. So, it is questionable whether we should describe the particles being further apart in a liquid, rather than just taking up a little more space.

In a molecule, the electron actually slots into spaces

Keith S. Taber

Mohammed was a participant in the Understanding Science Project. When interviewed in the first term of his upper secondary (GCSE) science course (in Y10), he told me he had been learning about ionic bonding in one of his science classes. Mohammed had quite a clear idea about ionic bonding, which he described in terms of the interactions of two atoms where "they both want to get full outer shells", leading to salt which was "like two atoms joined together":

The "two atoms joined together" sounds much like a molecule (and it is very common for students to identify molecule like ion-pairs even in representations of extensive ionic lattices), so I asked Mohammed about this:

Can I see these atoms?

No. They're really small. Because the wavelength of visible light is actually too like large to see the atoms, they just pass over them.

Okay, so I can't see them. But I can imagine them, can I?

Yeah.

So if I could imagine a sodium atom and chlorine atom, and then they form salt, what would it look like afterwards? How could I imagine it afterwards.

Oh it's like two atoms joined together.

That sounds like a molecule to me?

It's not actually, like, joined.

No?

Because I know that whenever things of opposite charge, I know two rods, when they come together, they don't actually touch, so they don't exactly touch, but they are very close, two atoms close to each other

So a molecule would be different to that in some way, would it?

Yeah, a molecule's actually bonded

So how that different?

I think in a molecule, the electron actually slots into spaces.

I see, and it doesn't do that in this case?

No.

So Mohammed thinks that the interaction between the ions will be due to their electrical charges, but, for him, this may not count as a bond, as the forces just hold the ions ("atoms") close together, and do not actually join them. Mohammed's idea of the atoms not actually touching, "they don't actually touch, so they don't exactly touch", is transferring a notion from the familiar world of macroscopic phenomena (where things touch, or they do not touch) to the submicroscopic world of quanticles that do not have definitive size/volume, and do not actually have distinct surfaces, so touching is a matter of degree. There is no more (or less) 'touching' in a covalent bond than in ionic bonding. So according to Mohammed the ions do not form a molecule, as in a molecule there would some kind of more direct joining – he suggests something like an interlocking with electrons from one atom slotting into spaces on another.

Interestingly, Mohammed bases his notion that the ions would not touch on a general principle that he considers to apply whenever considering things of opposite charge – which he justifies on his knowledge that "two [charged] rods, when they come together, they don't actually touch". He may be misremembering something here – or he may have seen a demonstration of suspended charged rods of the same material (so either both negatively or both positively changed) that when one is moved closer to the other the rods repel. Whatever the source, Mohammed seems to feel he has a valid general principle that he can apply here that act as a grounded learning impediment channelling his thinking about the case under discussion along 'the wrong lines'.

Mohammed's notion of the ionic bonding as being just due to forces rather than being a proper bond is very similar to a common alternative conceptions of ionic bonding which sees ions in a lattice only having a limited number of ionic bonds depending upon valency (the valency conjecture) but bonded with other coordination counter-ions by 'just forces' (the just forces conjecture) – although here Mohammed suspected that all ionic bonding fell short of being proper chemical bonds.

This is a very mechanical model of the covalent bond, whereas the scientific model presents bonding as more of a process than a material mechanical link. However teaching models often present bonding this way, and sometimes molecules are modelled in terms of jigsaws with atoms or radicals as pieces to be slotted together. Although such models are only meant to provide a simple analogy for the bonding they may act as learning impediments if learners take them too 'literally' as realistic representations and transfer inappropriate associations from the model to their understanding of the system being modelled.

Mohammed also uses similar language when asked about salt dissolving in water, as the charge of the water forces the sodium and chlorine ions to slot into certain places within the water molecules *.

In a sponge, the particles are spread out…

In a sponge, the particles are spread out more, so it can absorb more water 

Keith S. Taber

Morag was a participant in the Understanding Science Project. In her first term of secondary school, she told me that he had learnt about particles. Morag had explained, and simulated through role play for me, the arrangements of particles in the different states of matter (See: So if someone was stood here, we'd be a solid.) She had also emphasised just how tiny the particles were, "little, little-little-little things", and so how many there were in a small object: "millions and millions and millions". This suggested she had accepted and understood the gist of the scientific model of submicroscopic particles.

Yet as the conversation proceeded, Morag suggested the macroscopic behaviour of sponge in absorbing water could be explained by the arrangement of particles leaving space for the water. This is perhaps a reasonably, indeed quite imaginative, suggestion at one level, except that the material of a sponge is basically solid (where, as Morag recognised, that the particles would be very close together). A sponge as whole is more like a foam, with a great volume of space between the solid structure (where air can be displaced by liquid) and an extensive surface area.

Do you think it is important to know that everything is made of particles?

No.

It's not important?

Well it could be important, but it's not that important. Well, you see, like that [indicating the voice recorder used to record the interview] has got like lots and lots of particles pushed together this [Morag gestures]…But a sponge, the particles are like, the particles are more kind of like, they're still the same, but it's just spread out more, so it can absorb more water.

Oh I see, so are you saying that the same particles are in my little recorder, as in the sponge.

Yeah, they're the same, but there's just more of them in one than there would be in the other.

The failure here is perhaps less Morag's inappropriate explanation, than the tendency to teach about the ideals of solids, liquids and gases, which strictly apply only to single substances, where most real materials students come across in everyday life are actually mixtures or composites where the labels 'solid', liquid' and 'gas' are – at best – approximations.

Teaching has to simplify complex scientific ideas to make them accessible to students of different ages, so often teaching models are used. But sometimes simplifications can cause misunderstandings, and so the development of alternative conceptions. If 'everything is a solid, liquid or gas' is used as a kind of teaching model, or even presented as a slogan or motto for students to echo back to the teacher, when lots of things students come across in everyday life (e.g., butter, clouds, the pet cat – a bathroom sponge) do not easily fit these categories, and this is likely to lead to students overgeneralising.

Although it is often not possible to assign a single simple cause to a student's flawed thinking, this could be considered likely to be an example of a pedagogic learning impediment (a type of grounded learning impediment) in chemistry: a case where an approach to teaching can lead students' thinking in unhelpful directions.

Dissolving salt is a chemical change as you cannot turn it back

Dissolving salt is a chemical change as you cannot turn it back as it was before

Keith S. Taber

Sandra was a participant in the Understanding Science Project. When I interviewed Sandra about her science lessons in Y7 she told me "I've done changing state, burning, and we're doing electricity at the moment". I asked her about burning:

Well, tell me a bit about burning then. What's burning then?
It's just when something gets set on fire, and turns into ash, or – has a chemical change, whatever.
Has a chemical change: what's a chemical change?
It means something has changed into something else and you can't turn it back.
Oh I see. So burning would be an example of that.
Yeah.

So far this seemed to fit 'target knowledge'. However, Sandra suggested that dissolving would also be a chemical change. Dissolving is not normally considered a chemical change in school science, but a physical change, the distinction is a questionable teaching model. (Chemical change is said to involve bond breaking/making, and of course dissolving a salt does involve breaking up the ionic bonding to form solvent-solute interactions.)

Are there other examples?
Erm – dissolving.
So give me an example of something you might dissolve?
Salt.
Okay, and if you dissolve salt, you can't get it back?
Not really, not as it was before.
No. Can you get it back at all?
Sort of, you can like, erm, make the, boil the water so it turns into gas, and then you have salt, salt, salt on the, left there. Sometimes.
But you think that might not be quite the same as it was before?
No.
No. Different in some way?
Yeah
How might it be different?
Be much smaller.
Oh I see, so do you think you'd have less salt than you started with?
You'd have the same, but there would just be more particles, but they'd be smaller.
Ah, so instead of having quite large grains you might have lots of small grains
Yeah.

So Sandra was clear that one could dissolve salt, and then reclaim the same amount of salt by removing the solvent (water) which from the canonical perspective would mean the change was reversible – a criterion of a physical change.

Yet Sandra also thought that although the amount of salt would be conserved, the salt would be in a different form – it would have different grain size. (Indeed, if the water was boiled off, rather than left to evaporate, it might indeed be produced as very small crystals.)

So, Sandra seemed to have a fairly good understanding of the process, but because of the way she interpreted the criterion of a chemical change, something [salt] has changed into something else [solution] and you can't turn it back [with the same granularity]. Large grains will have changed into small grains – so this would, to Sandra's mind, be a chemical change.

Science teachers deserve a great deal of public appreciation. A teacher can teach something so that a student learns it well – and yet still form an alternative conception – here because of the inherent ambiguity in the ways language is used and understood. Sandra's interpretation – if you start off with large particles and end up with smaller particles then you have not turned it back – was a reasonable interpretation of what she had learnt. (It also transpired there was ambiguity in quite what was meant by particles.)