Keywords

These keywords were added by machine and not by the authors. This process is experimental and the keywords may be updated as the learning algorithm improves.

David Gross has described the early 1960s as a period of “experimental supremacy” [28, p. 9099]. The theoretical situation was almost entirely phenomenologically-oriented, with a profusion of new particle data being generated by experiments at Brookhaven, CERN, DESY, SLAC, and elsewhere. Theory was in a rather sorry state. Most of the work was concerned with model building to try and get some kind of foothold on the diversity of new phenomena coming out of the latest generation of particle accelerators. There was genuine uncertainty about the correct framework for describing elementary particles, and even doubts as to whether there were such things as elementary particles. Footnote 1

One of the central problems was triggered by the strong interactions, involving hadrons,Footnote 2 describing the properties of nuclei. True to their name the strongly interacting particles have large coupling constants determining how strongly they interact with one another, so the standard field theoretical tool of expanding quantities in powers of these constants fails to give sensible results.Footnote 3 Steven Weinberg notes that the “uselessness of the field theory of strong interactions led in the early 1950s to a widespread disenchantment with quantum field theory” [55, p. 17]. It didn’t end with the strong interactions: the weak interaction too was then described by the non-renormalizable Fermi theory. This situation led to a move to bypass quantum field theory and instead deal directly with the fundamental constraints—and other general properties characteristic of the strong interaction—on the S-matrix that are expected of a good relativistic quantum theory (i.e. the scattering probability amplitude).Footnote 4 As Pierre Ramond writes, “[i]n the absence of a theory, reliance on general principles alone was called for” [46, p. 503].

String theory did not spontaneously emerge from a theoretical vacuum; it emerged precisely from the conditions supplied by this profound foundational disenchantment.Footnote 5 With hindsight, the earliest developments in string theory—i.e. the dual resonance models alluded to in the previous chapter—can be viewed as perfectly rational and progressive steps given the state of physics just prior to it. In this first part we describe this state of affairs, and introduce the mathematical and physical concepts, formalism, and terminology necessary in order to make sense of early (and large portions of later) string theory.Footnote 6 However, many of the old concepts still make an appearance in modern string theory despite being not so well-known. This part might therefore also serve as a useful primer for those wishing to learn some string theory by providing some of the original physical intuitions and motivations.

1 Hadrontology

Recall that hadrons come in two families: baryons (particles composed of three quarks, such as the protons and neutrons making up an atomic nucleus) and mesons (force-mediating particles composed of two quarks, (a quark and an anti-quark), such as the pions and kaons found in cosmic rays)—particles that do not interact via the strong force are called leptons. The interactions amongst the components of nuclei were originally thought to be mediated entirely by \(\pi \)-mesons (a name contracted to ‘pions’). However, the early models did not consider hadrons as internally structured entities composed of point-like constituents that interact through hard collisions, but as extended objects with ‘soft’ interactions.Footnote 7 From our present position we would say that the models were latching on to low-energy, long-range aspects of hadron physics in which the pions were the tip of an iceberg. There were many more mesons lurking below the surface. Unifying the profusion of mesons and baryons posed one of the most serious challenges of mid-twentieth century physics.

The challenge was further intensified as technological advances made possible proton acceleratorsFootnote 8 and bubble chambers capable of registering events involving hadrons by photographing bubbles formed by charged particles as they dart through a superheated liquid, thereby superseding earlier cosmic rays observations.Footnote 9

Of course, quantum mechanics renders events probabilistic. This infects the natural observables in particle physics too. One of the observable quantities is the scattering cross-section (which basically offers a measure of the scattering angle made by colliding beams, or a beam and a static target). This tells you the likelihood of a collision given that two particles are moving towards one another. The magnitude of the cross-section is directly proportional to this likelihood. The cross-section itself is a function of the energy of the incoming beams, and if one examines the behaviour of the cross-section as a function of this energy, one can find peaks such that one can ask whether they correspond to particles or not.

Analysing the data from these scattering experiments pointed to the production of very many more new particles, or ‘hadronic resonances’ (very short lived, ‘fleeting’ strongly interacting particlesFootnote 10 corresponding to sharp peaks in the total cross section, as a function of the energy)—of course, the strong interaction’s being strong implies that such particle production will be plentiful. As described in the Particle Data Group’s documents, resonant cross sections are described using the Breit-Wigner formula:

$$\begin{aligned} \sigma (E) = \frac{2J + 1}{(2S_{1}+1)(2S_{2}+1)} \frac{4\pi }{k^{2}} \displaystyle \left[ \frac{\varGamma ^{2}/4}{(E - E_{0})^{2} + \varGamma ^{2}/4} \displaystyle \right] B_{in}B_{out} \end{aligned}$$
(2.1)

where \(E\) is the energy in the centre of mass frame, \(J\) is the resonance spin, \((2S_{1}+1)\) and \((2S_{2}+1)\) are the polarisation states of a pair of incident particles, \(k\) is the initial momentum in the centre of mass frame, \(E_{0}\) is the resonance energy (again in the centre of mass frame), \(\varGamma \) describes the resonance width (with \(\frac{1}{\varGamma }\) giving the mean resonance lifetime), and the \(B_{in}B_{out}\) pair describe the resonance branching fractions for the incoming and outgoing channels, where \(B_{channel}\) would be computed as \(\frac{\varGamma _{channel}}{\varGamma _{all}}\) (that is, one counts the total number of decays through some channel relative to the total number of particles produced)—see http://pdg.lbl.gov/2013/reviews/rpp2012-rev-cross-section-formulae.pdf for more details.

Fig. 2.1
figure 1

Particle tracks showing the annihilation event of an anti-proton within a liquid hydrogen bubble chamber (using the PS coupled to the 80 cm Saclay chamber used by CERN—image taken in 1961). Decay products are a negative kaon, a neutral kaon, and a positive pion. Image source CERN, 1971

Fig. 2.2
figure 2

The associated reactions of the previous photograph. Image source CERN annual report of 1961, [3, p. 93]

The search for patterns in this jumble of data led to the discovery of a new symmetry principle and a deeper quark structure underlying the dynamics of hadrons. This work can be viewed in terms of a drive to systematise.Footnote 11 A central concern was whether these new particles (or, indeed any of the particles) were ‘fundamental’ (i.e. elementary)—with the sheer number of different particle types naturally casting doubt on the idea that they were all elementary. If so, then the others might be constructed as bound states of some small number of elementary particles.Footnote 12

One of the most hotly pursued approaches, S-matrix theory, involved focusing squarely on just those properties of the scattering process—or more precisely of the probability amplitude for such a scattering event—that had to be obeyed by a physically reasonable relativistic quantum field theory. The combination of these general principles with (minimal) empirical evidence drawn from observations of hadrons was believed to offer a way of (eventually) getting a predictive physics of strong interactions.Footnote 13 In its most radical form, espoused by Berkeley physicist Geoffrey Chew, the question of which hadrons were elementary and which were bound states was simply not appropriate; instead, one should treat them democratically, as on all fours.Footnote 14

The S-matrix was originally developed by John Wheeler, as a way of condensing the complex bundle of information that is needed to describe a collision process, encapsulating the experimentally accessible information about any scattering experiment one could think of. Heisenberg actually named the object that achieves this condensation and imbued it with far more significance than Wheeler ever did.Footnote 15 Wheeler saw it as a mere tool “to deal with processes to be analysed by a more fundamental treatment” [56]. This might, as in the case of quantum electrodynamics [QED] be provided by a quantum field theory, which delivers up an S-matrix as an infinite expansion in the coupling constant (as we saw, in the case of QED this is the fine-structure constant \(\alpha _{\textit{EM}} = \frac{e^{2}}{4\pi }\)).Footnote 16 Alternatively, one can sidestep talk of fields entirely, and focus on the scattering probability amplitude itself, which after all should contain all physically observable information (including the cross sections mentioned above, which can be written in terms of the matrix elements).

In this latter sense the S-matrix has an affinity with Bohr’s positivistic strategy of ignoring what happens between energy transition processes involving electrons orbiting atoms. In this case what is ignored (as unphysical or meaningless since unobservable, since too short-lived) are the unmeasurable processes occurring between initial and final states of a collision process.Footnote 17 Rather than describing what happens at the precise spacetime point (the vertex) at which the two or more particles meet (in which case there is no measurement to ascertain what is happening), one focuses on the measurable ‘free’ (non-interacting) situation when the systems were not and are no longer able to causally interact (mathematically speaking, at infinity, in the asymptotic limits), and therefore the particles have straight trajectories at constant velocities. In effect one draws a black box around the innards of the process and focuses on the particles entering and leaving the box and the probabilities of their doing so. This is somewhat paradoxical since the interaction between particles is described by an expression involving the particles’ being far apart!

The S-matrix catalogues these possible relations between inputs and outputs along with their various probabilities. Measurable quantities such as scattering cross-sections can be written in terms of the matrix elements of the (unitary) S-matrix operator \(\mathsf {S}\). Recall that in quantum mechanics the state of a system is represented by a wave function \(\psi (p)\), a square-integrable function of the system’s momentum \(p\) (a 3-vector). For \(n\) particles it is a function of all the particles’ momenta \(p_{1}, ..., p_{n}\) (each a 3-vector).Footnote 18 The S-matrix is then an operation that transforms an initial state (a free wavefunction) of such incident particles to a final state (another free wavefunction), which, under the action of the unitary operator \(\mathsf {S}\), will have the general form of a superposition of all possible final states. The amplitude for finding one of these final states (say \(|p'_{1}, p'_{2} \rangle \)) in a measurement (for which the initial state is \(|p_{1}, p_{2}\rangle \)), is given by \(\langle p'_{1}, p'_{2} | \mathsf {S} | p_{1}, p_{2}\rangle \):

figure a

As in many episodes in the history of physics, what was essentially a mathematical result, here from complex analysis, led in 1959 to a breakthrough in physical theory. Analytic continuation allows one to extend the domain of definition of a complex function. A (complex) function is said to be analytic (or holomorphic, in mathematical terms) if it is differentiable at every point in some region. It was already known, thanks to the work of Gell-Mann, Chew, and others, that the S-matrix was an analytic function of its variables (representing physical quantities: the momenta of ‘incoming’ and ‘outgoing’ particles). This allowed the properties of the S-matrix to be probed almost independently of field theoretical notions in a quasi-axiomatic fashion (with very little by way of direct experimental input). The S-matrix theory (also known as the ‘theory of dispersion relations’,Footnote 19 though the links between dispersion relations and Heisenberg’s theory took more time to emerge) then sought to derive the S-matrix by imposing various natural consistency conditions on it: Lorentz invariance, crossing,Footnote 20 unitarity, and analyticity (see the box below).

  • Lorentz invariance is satisfied when physical quantities are unchanged by Lorentz transformations (of the form \(x'^{\mu } = \varLambda ^{\mu }_{\nu } x^{\nu }\) for all 4-vectors \(x^{\nu } = (x^{0}, \mathbf {x}) = ( {t} , \mathbf {x})\) and Lorentz tensors \(\varLambda ^{\mu }_{\nu }\)). (Of course, this also implies that energy, momentum, and angular momentum are conserved.)

  • Analyticity is satisfied just in case a scattering amplitude \(A\) is an analytic function of the Lorentz invariant objects used to represent the physical process in which one is interested. This formal condition is the mathematical counterpart of causality (i.e. the outlawing of effects preceding causes). (This condition has its origins in the dispersion relations of classical optics—see footnote 19.)

  • Crossing is a symmetry relating a pair of processes differing by the exchange of one of the input and output particles (mapping particle to anti-particle and vice versa); for example, \(a+b \rightarrow c+d\) and \(a + \overline{c} \rightarrow \overline{b}+d\) (where \( \overline{b}\) and \( \overline{c}\) are \(b\) and \(c\)’s anti-particles).

  • Unitarity is simply the condition that the scattering matrix \(S\) is unitary: \(S^{\dag }S = 1\). Or, in other words, probability (that is, the squared modulus of the amplitude) must be conserved over time. (This also includes the condition of coherent superposition for reaction amplitudes.)

As indicated above, one of the central objects of the physics of elementary particle physics is the scattering (or transition) amplitude \(A\). This is a function that churns out probabilities for the outcomes of collision experiments performed on pairs of particlesFootnote 21—note, this is not the same as the matrix of such described above. It takes properties of the particles as its argument. For example, the function might depend on the energy \(E\) of the collision event and the scattering angle \(\theta \) representing a particle’s deflection \(f(E,\theta )\) thus encodes the nature of this interaction. The general representation involves the incoming energy and the momentum that is transferred in the collision, \(s\) and \(t\) respectively, defined as follows:

  • \(t\) is the square of the difference between the initial and final momenta of the particles involved in some process (also known as “the momentum transfer”):

    $$\begin{aligned} t = (p_{a} - p_{c})^{2} = (p_{b} - p_{d})^{2} \end{aligned}$$
    (2.2)
  • \(s\) is the square of the sum of the momenta of the initial states on the one hand and the final states on the other:

    $$\begin{aligned} s = (p_{a} + p_{b})^{2} = (p_{c} + p_{d})^{2} \end{aligned}$$
    (2.3)

We denote the incoming momenta of the particles, \(p_{a}\) and \(p_{b}\), with outgoing momenta \(-p_{c}\) and \(-p_{d}\). In this process there is a conservation of total momentum (4-momentum); i.e. \(p_{a} + p_{b} = p_{c} + p_{d}\) (also, \(p_{i}^{2} = m_{i}^{2}\), with \(m_{i}\) being the \(i\)th particle’s mass).Footnote 22 The scattering amplitude is, then, a function of certain conserved (invariant) quantities (‘channel invariants’). Suppose we have some process involving a pair of incoming particles going into some pair of outgoing particles (of the same mass \(m\), for simplicity): \(a+b \rightarrow c+d\). This will involve a 4-point amplitude \(A(s,t)\). The amplitude is then written as \(A(s,t) \sim \beta (t) (s/s_{0})^{\alpha (t)}\) (where \(\beta \) is a residue function). The squared modulus of this object delivers the observable scattering cross-section discussed above.

The Mandelstam variables define reaction channels as follows (see Fig. 2.4):

  • The reaction \(a + b \rightarrow c+d\) occurs in the \(s\)-channel, with the physical (real) region defined by values \(s \ge (m_{a} + m_{b})^{2}\).

  • The ‘crossed’ reaction \(a + \overline{c} \rightarrow \overline{b}+d\) occurs in the \(t\)-channel (as noted in the box above), with the physical (real) region defined by values \(t \ge (m_{a} + m_{c})^{2}\).Footnote 23

Fig. 2.3
figure 3

Graph showing the number of papers published on S-matrix theory (or the S-matrix) following Heisenberg’s paper in 1943, with a significant growth occurring in the 1960s. In his survey of models for high-energy processes, John Jackson found that, between 1968 and the first half of 1969, “various aspects of S-matrix theory, with its ideas of analyticity, crossing and unitarity, accounted for 35 % of the theoretical publications” [30, p. 13]; cf. [52, p. 285]. Image source Thompson-Reuters, Web of Science

Recall that Feynman diagrams were originally intended to provide a mathematical representation of the various contributions to the S-matrix in the context of perturbative (Lagrangian) field theories. However, in the late 1950s Landau [34] had instigated the examination of the links between Feynman graphs and singularities of the S-matrix, thus liberating the former from weakly-coupled quantum field theories to which they were previously thought to be hitched. The singularity conditions that Landau found pointed to a correspondence between tree graphsFootnote 24 and poles (and loop diagrams and branch points). Thus was born the idea that general conditions imposed on the structure of the scattering amplitude might be enough to determine the physical behaviour of particles.

These considerations led to a variety of features that could be aimed at in model building. It was from this search that the Veneziano model was born. Before we discuss that model, we first need to say something about some important intervening work, of Tullio Regge, Stanley Mandelstam, and Geoffrey Chew, that will help us make better sense of the foregoing.

2 Chew’s Boots and Their Reggean Roots

In 1959 Tullio Regge [47] suggested that one think of solutions to the Schrödinger equation for the potential scattering problem in terms of the complex plane, using complex angular momentum variables (which, of course, take on discrete values). This ignited a surge of research in linking ‘Regge theory’ to the world of hadrons and high energy (special relativistic) physics.Footnote 25

A singularity of a complex function (i.e. a point where the value of the function is zero or infinity for some argument) is known as a pole (a tree graph in graphical terms, with loops corresponding to branch points). Regge focused on the potential scattering problem, where the amplitudes become simple poles in angular momentum (i.e. at certain special values of the momenta). The locations of these poles is determined by the energy of the system and the poles themselves were taken to correspond to the propagation of intermediate particles. As one tunes the energy parameter, one gets a graph (a Regge trajectory) describing the properties of resonances and scattering amplitudes (for which the transfer of momentum is large). In the relativistic case one must introduce another class of singularity in angular momentum, in particular at \(j = -1\). Stanley Mandelstam tamed these singularities by introducing a second Riemannian hyperplaneFootnote 26 of the complex \(j\)-plane and performing branch cuts in the \(j\)-plane, known as “Regge cuts”.Footnote 27

A Regge pole is then the name given to a singularity that arises when one treats angular momentum \(J\) as a complex variable.Footnote 28 Physically a Regge pole corresponds to a kind of particle that ‘lives’ in the complex angular momentum plane, whose spin is linearly related to its mass. Tuning the energy of such a particle to a value which would spit out an integer or half-integer value for the spin would produce a particle that one ought to be able to detect. Confirmation of this relationship was indeed found in early hadron spectroscopy which generated Regge plots showing (for mass squared plotted against spin) a linearly rising family of particles on what became known as a ‘Regge trajectory’ (see Fig. 2.5).Footnote 29 In this way specific types of particles could be classified by these trajectories, each trajectory containing a family of resonances differing with respect to spin (but sharing all other quantum numbers).

There was a curious feature about some of the spin values,Footnote 30 as represented in the plots of Regge trajectories, namely that they were seemingly unbounded from above. Particles with large spins are more like finite-sized objects possessing angular momentum (from real rotationFootnote 31). In the case of baryons, one can find experimentally observed examples of spin \(J\) \(=10\)! According to Regge theory, the high energy behaviour of scattering amplitudes is dominated by the leading singularity in the angular momentum Argand plane. Crucially, if such a singularity is a pole at \(J=\alpha (t)\) (in other words, a Regge pole) then the scattering amplitude has the asymptotic behaviour: \(\varGamma (1-\alpha (t)) (1+ e^{-i\pi \alpha (t)}) s^{\alpha (t)}\) (where \(s\rightarrow \infty \) and \(t<0\)).

Fig. 2.4
figure 4

The Mandelstam diagram providing a representation of \(A(s,t,u)\) in terms of double spectral functions, \(\rho _{12}\), \(\rho _{31}\), and \(\rho _{23}\), which are zero except in the shaded region (corresponding to values above the intermediate-state threshold). Image source [12, p. 120]

The bootstrap approach grew out of these developments of Regge and Mandelstam.Footnote 32 In dispersion theory one tries to generate physics from a few basic axioms, such as Lorentz invariance, unitarity, and causality discussed above. These are used as (high-level physical) constraints on the space of possible theories as input data from the world is fed in. The dispersion theory approach and the old S-matrix approach were merged together in Chew’s ‘bootstrap’ approach to physics.Footnote 33

A crucial component of Chew’s approach was the ‘pole-particle’ correspondence. According to this principle, there is a one-to-one correspondence holding between the poles of an (analytic) S-matrix and resonances, so that the position of a pole in the complex energy plane gives the mass of the resonance while the residue gives the couplings. When the pole is complex, the imaginary part gives its lifetime. The idea was that the axioms of the dispersion approach would uniquely pin down the correct S-matrix, and thereby deliver physical predictions. The focus would be on the analytic properties of the S-matrix. The theory had some degree of success at a phenomenological level.

Presently, of course, our best description of nature at very small subatomic scales is couched in the framework of quantum field theory [QFT]—a framework Chew believed unhealthily imported concepts from classical electromagnetism. It is thought that there are six fundamental leptons and six fundamental quarks. These are bound together by forces that are understood as involving quantum fields. The unified theory of the weak and electromagnetic interactions, the electroweak force, is understood via the exchange of four kinds of particle: the photon, the \(W^{+}\), the \(W^{-}\), and the \(Z^{0}\). The strong force is mediated via the exchange of eight types of massless gluon. The standard model also involves Higgs particles, \(H^{0}\), whose associated field is responsible for the generation of the masses of observed particles.Footnote 34 In quantum field theory the dynamics is delivered through a Lagrangian, from which one derives equations of motion. Essentially what Chew proposed was to eliminate equations of motion in favour of general principles. In the case of strong interactions, at least, Chew believed that a Lagrangian model simply wasn’t capable of delivering up a satisfactory S-matrix.

At the root of Chew’s proposal was the belief that field theory could simply not cope with the demands imposed by strong interaction physics. He wrote that “no aspect of strong interactions has been clarified by the field concept” [6, p. 1]. Though there was a family of hadrons, no family members appeared to be fundamental, and a field for each and every hadron would result in filling space with an absurd number of fields. For this reason, Chew suggested that all hadrons should be treated on an equal footing: neither more nor less fundamental than any other. The notion of fundamentality dropped out in favour of nuclear democracy, with the particles understood as in some sense composed out of each other as in footnote 33, with the forces and particles bundled together as a package deal. Chew expresses it as follows:

The forces producing a certain reaction are due to the intermediate states that occur in the two “crossed” reactions belonging to the same diagram. The range of a given part of the force is determined by the mass of the intermediate state producing it, and the strength of the force by the matrix elements connecting that state to the initial and final states of the crossed reaction. By considering all three channels [i.e., orientations of the Feynman diagram] on this basis we have a self-determining situation. One channel provides forces for the other two—which in turn generate the first [6, p. 32].

Fig. 2.5
figure 5

A Regge trajectory function \(\alpha (t)\) representing a rotational sequence of states (of mesons) of ever higher spins. The relationship with resonances (and bound states) comes about from the fact that when \(\alpha (t)\) is a positive integer for some value of the argument \(t\), then a bound state or resonance exists at that \(t\)-value, with spin read off the horizontal. For example, in this picture we have at \(t=3\) the resonance \(\alpha (3) = 3\) and at \(t=-1\) the bound state \(\alpha (-1) = 0\). The various states given in this way generate a family: the Regge trajectory. A horizontal trajectory, \(\alpha (t) = const.\), would represent particles of constant spin (elementary particles), while a non-zero slope represents particles of varying spin (composite particles). Image source [15, p. 1]

A further development that played a crucial role was made by Chew’s postdoc student at Berkeley, Stanley Mandelstam. He had discovered a way to resolve a problem in understanding the strong interaction in terms of particle exchange (à la YukawaFootnote 35). The problem was that the hadrons were short range, and therefore massive—Yukawa had calculated a characteristic mass of \(100\) MeV, corresponding to a sub-nuclear range of the strong force of \(10^{-13}\) cm. The old cosmic ray observations delivered a candidate for such a particle in the form of the pion. Yet, by the late 1950s, particles were also being discovered with spins greater than 1, increasing linearly. This would imply that the exchange forces would also grow in such a way, without limit. Referring back to the discussion above, this would further imply that the scattering cross-section describing the size of the area over which the particles interact would also grow indefinitely. This is in direct conflict with the idea that exchanging massive particles demands smaller areas: the more massive the particles are, the less capable they are of covering large distances.

The solution was to treat the entire series of particles (with increasing spins) laid out along a Regge trajectory as the subjects of exchange (named a “pomeron” by Vladimir Gribov, after Pomeranchuk)—that is, rather than the individual points lying within the trajectories.Footnote 36 Applying this procedure keeps the cross-sections finite—a calculation that was performed by Chew and Steven Frautschi [5].Footnote 37

3 Enter Duality

An important step in the bootstrap approach was the principle of duality introduced by Dolen, Horn, and Schmid in 1967, at Caltech (they referred to it as “average duality” or “FESR duality”, for reasons given below).Footnote 38 They noticed that Regge pole exchange (at high energy) and resonance (at low energy) descriptions offer multiple representations (or rather approximations) of one and the same physically observable process. In other words, the physical situation (the scattering amplitude, \(A(s,t)\) Footnote 39) can be described using two apparently distinct notions (see Fig. 2.6):

Fig. 2.6
figure 6

Graphical representations of two descriptions of the hadronic scattering amplitude: In the left diagram one has resonance production (with \(\pi \) and \(N\) colliding to generate \(N^{*}\), which decays after a short time back into \(\pi \) and \(N\)); on the right hand side one has Regge pole exchange (i.e. an interaction in which \(\pi ^{-}\) and \(p\) exchange a \(\rho \)-meson, transforming quantum numbers to become \(\pi ^{0}\) and \(n\)). Image source C. Schmid [49, p. 257]

  • A large number of resonances (poles) exchanged in the s-channel.

  • Regge asymptotics: \(A(s,t)_{s \rightarrow \infty } \sim \alpha (s)^{\alpha (t) - 1}\), involving the exchange of Regge poles in the t-channel.

That these are in some sense ‘equivalent’ in terms of the physical description was elevated to a duality principle Footnote 40:

DHS Duality Direct \(s\)-channel resonance particles are generated by an exchange of particles in the \(t\)-channel.

This has the effect that the representative Feynman diagrams for such processes are identified to avoid surplus states, known as “double counting”. For this reason, the two contributions to the amplitude are not to be summed together: summing over one channel is sufficient to cover the behaviour encapsulated in the other. This was matched by the experimental data. So-called “interference models” would demand that the two descriptions (both \(s\)- and \(t\)-channel contributions) be added together like ordinary Feynman tree diagrams, which would be empirically inadequate of course (see Fig. 2.7). As with any duality there is an associated epistemic gain: if we know about the resonances at low energies, we know about the Regge poles at high energies.Footnote 41

Fig. 2.7
figure 7

Different methods of computing amplitudes with the interference model (top), associated with a picture of elementary particles, and the dual resonance model (bottom), associated with composite entities. In the former one sums over the contributions from both channels, while the latter identifies them in accordance with the principle of duality. Image source [39, p. 265]

One can make some physical sense of the existence of such a duality by thinking about the ‘black box’ nature of the scattering methodology, as discussed previously. Since one makes measurements only of the free states (the asymptotic wave-functions), one cannot discern the internal structure between these measurements, and so given that both the \(s\)-channel (resonance) and \(t\)-channel (interaction via exchange) situations have the same asymptotic behaviour, they correspond to ‘the same physics’. However, the precise mathematical reason would have to wait first for the formulation of a dual amplitude, and then for the string picture, at which point it would become clear that conformal invariance was grounding the equivalence between such dual descriptions.

Mention must be made of the Finite Energy Sum Rules (i.e. where the energy has been truncated or cut in \(s\)), which are further consistency conditions, flowing from analyticity.Footnote 42 They are an expression of a linear relationship between the particle in the \(s\)- and \(t\)-channels and were a crucial step on the way to the DHS duality principle. They have enormous utility in terms of applications, not least in allowing the low and high energy domains of scattering amplitudes to be analytically connected: at high energies the scattering amplitude will be ruled by a handful of Regge poles (in the so-called ‘crossed’ \(t\)-channel) viewed at low energies the amplitude will be ruled instead by a handful of resonances (in the so-called ‘direct’ \(s\)-channel), as above. Thus, the FESR already establish a kind of duality between these two regimes so that \(t\)-channel (Regge) values can be determined from \(s\)-channel resonances. More formally, one begins with the (imaginary part of the) low energy amplitude characterised by resonances (which sits on the left hand side of the FESR equation) and builds up the Regge terms by analytic continuation (cf. [49, p. 246]). Schematically one has (borrowing from [43, p. 204]):

$$\begin{aligned} \langle \text {Im} f (\text {Resonance}) \rangle = \langle \mathrm {Im} f (\text {Regge}) \rangle \end{aligned}$$
(2.4)

The averaging refers to the fact that one is integrating over Regge and resonance terms (Fig. 2.8).Footnote 43 The FESR are formally expressed as follows:

$$\begin{aligned} \mathbf {FESR:} \;\; \int _{o}^{N} \mathrm {Im} \, A^{(-)} ({\textit{v}},t) dv = \sum _{i} \beta _{i}(t) \frac{N^{\alpha _{i}(t) + 1}}{\alpha _{i}(t) + 1} \end{aligned}$$
(2.5)

Hence, DHS duality is sometimes also called FESR-duality.

Fig. 2.8
figure 8

A plot indicating duality for the \(\pi N\)-amplitude, \(A^{'(-)}\). One can see that \(2 Im A^{'(-)}\) has large fluctuations at low energy values, but latches on to the \(\rho \) Regge term on the average. Image source C. Schmid [49, p. 260]

Though this duality in some ways embodies Chew’s Nuclear Democracy (since, in the case of \(\pi \pi \) scattering, both channels contain the same particles) it also paved the way for a departure from this picture. Using diagrammatic representations of the duality, Harari and Rosner reinterpreted the duality in terms of the flow of hadron constituents (quarks and anti-quarks Footnote 44) and the exchange of such.

Although the link wasn’t explicitly made at the time, these diagrams, in eliminating the links and vertices from standard Feynman graphs, already contain the germ of what would become string scattering diagrams according to which only the topological characteristics are relevant in the scattering process—one can easily see that the exchange and resonance diagrams are deformable and so topologically equivalent. This equivalence was given a graphical representation in the work of Haim Harari (see Fig. 2.9).

Fig. 2.9
figure 9

Haim Harari’s duality diagram for a multi-particle process. The top diagram amounts to an equivalence class of the diagrams beneath it, in the sense that any of the five ordinary Feynman graphs provides complete information about the amplitude. Image source [9, p. 563]

Harari was then working at the Weizmann Institute. At around the same time, at Tel-Aviv University, Jonathan Rosner also came up with the idea of duality diagrams.Footnote 45 Rosner’s version can be seen in Fig. 2.10.

Fig. 2.10
figure 10

Jonathan Rosner’s graphical representation of duality. A graph will exhibit duality in only those channels in which it is planar (no crossed quark lines). Mesons are quark/anti-quark pairs, \(q\overline{q}\) and baryons are triplets of quarks \(qqq\). Here, a is planar in the \(s\) and \(t\) channels, with an imaginary part at high \(s\)—this represents baryons in \(s\) and mesons in \(t\). Duality implies that intermediate baryon states build an imaginary part at high \(s\). Graph b is planar only in \(u\) and \(t\), with no imaginary part at high \(s\). Image source [20, p. 689]

Since it makes an appearance in the following pair of chapters, we should also say something about the Pomeron (that is, the Pomeranchuk pole) in this context. The duality principle links Regge poles to resonances, but the Pomeron, with vacuum quantum numbers, falls outside of this scheme. It satisfies duality in a sense, but it turns out to be dual to the non-resonating background terms.

Another problematic issue was simple one pion exchange. The problem with this case, vis-a-vis duality, is that the amplitudes for such exchange processes are real-valued, whereas, as we have seen, duality involves only the imaginary parts of amplitudes. Though this problem was discussed (see, e.g., the remark of Harari following Chan’s talk at a symposium on Duality-Reggeons and Resonances in Elementary Particle Processes, [11, p. 399], it doesn’t seem to have been satisfactorily resolved until John Schwarz and André Neveu’s dual pion model in 1971.

As we will see in the next chapter, Veneziano’s achievement was to display a solution to FESR by the Euler Beta function (thus giving an implementation of a dual version of the bootstrap). The solution is an amplitude that displays precisely the Regge behaviour (that is, Regge asymptotics) and satisfies all of the principles laid out by the S-matrix philosophy (Lorentz invariance, analyticity, crossing, duality), apart from unitarity, on account of the particular approximation scheme employed (on which more later). The hope was that using the bootstrap principle, this framework could then eventually be employed to predict specific physical properties of hadrons, such as masses.

The ability of dual models to encompass so many, then ill-understood, features of hadronic physics led to their very quick take up. Quite simply, there was no alternative capable of doing what dual models did. Hence, though it was not then able to make novel testable predictions, even at this stage, the fact that it resolved so many thorny problems with hadrons, and explained so many features in a unified manner meant that it was still considered to be serious physics—though, it has to be said, not all were enamoured, precisely on the grounds that it failed to make experimental predictions.

Before we shift to consider the Veneziano model, a further important step towards the dual models, and away from Chew-style bootstrap models, was the introduction of the narrow-resonance (or zero-width) approximation alluded to above, which initially ignored the instability of hadrons, treating all of them instead as stable particles, with scattering and decays then progressively added as perturbations.Footnote 46 Stanley Mandelstam [38, p. 1539], wishing to model the rising Regge trajectories within the double dispersion relations approach, introduced the “simplifying assumption” that the scattering amplitude is dominated by narrow resonances (where the amplitude is understood to be approximated by a finite number of Regge poles). In this scheme, Mandelstam was able to implement crossing symmetry using the FESR. To achieve the rising, Mandelstam uses two subtraction constants,Footnote 47 which in turn generates a pair of new parameters into the scheme: the Regge slope \(a\) and the intercept \(b\) (now written, \(\alpha \) and \(\alpha (0)\) respectively). These two parameters are absolutely central to the physical implications of the early attempts to construct dual symmetric, Regge behaved models, and still play a vital role today. Mandelstam makes an additional (well-motivated) assumption that the trajectories built from these parameters, namely \(\alpha (s) = as+b\), do not rise “more than linearly with \(s\)” (p. 1542). For this reason, it might be prudent to call \(\alpha (s)\) the ‘Regge-Mandelstam slope’ rather than the Regge-slope.Footnote 48

4 A Note on Early Research Networks

For reasons that should by now be clear, those working on the S-matrix programme and the bootstrap approach to strong interaction physics play a ‘statistically significant’ role in string theory’s early life, the latter being an outgrowth of the former via the dual resonance model (as we will see in the subsequent pair of chapters). An important subset of the current string theory researcher network can be traced back quite easily to a small group of physicists from this period in the 1960s, all working in and around the S-matrix programme (or dispersion relations) and Regge theory. This is quite natural, of course, since the dual resonance models can be viewed as a culmination of the bootstrap approach (recall Cushing’s remark about superstring theory constituting “the ultimate bootstrap” [12]). The lines of influence are presented below.Footnote 49

figure b

In particular, we can see a clear clustering around Geoffrey Chew and Berkeley. In a key move, Chew invited Stanley Mandelstam over to Berkeley, as a postdoc, who brought over the skills of complex analysis. It seems that Chew liked to be in close proximity to his students, and held weekly group meetings with them to discuss what they were working on. This close proximity clearly led to Chew’s idiosyncratic positions being transmitted throughout the group.Footnote 50 Note that prior to joining Berkeley, Chew was based at the University of Illinois, Urbana-Champaign, together with Francis Low. Nearby, at the University of Chicago, were Nambu and Goldberger. Richard Brower, whom we will encounter later, had been at Berkeley, interacting closely with Chew and Mandelstam (his supervisor).

Note that John Schwarz was working on sum rules while at Princeton University in 1967. Schwarz’s advisor was Geoff Chew. While at Berkeley, heavily influenced by Chew, he would have been steered away from work on elementary quarks Footnote 51 and quantum fields. Of course, this can’t provide any explanation of why Schwarz and a few others from Chew’s workshop continued to avoid quantum field theory. After all, David Gross (one of the few responsible for laying the finishing touches to QCD) was also a student of Chew’s at roughly the same time as Schwarz and, indeed, the two shared an office during their three final years (1963–6), writing a joint paper in 1965.Footnote 52

Gross pinpoints the moment he became disillusioned with his supervisor’s approach following a remark from Francis Low, at the 1966 Rochester meeting:

I believe that when you find that the particles that are there in S-matrix theory, with crossing matrices and all of the formalism, satisfy all these conditions, all you are doing is showing that the S-matrix is consistent with the way the world is; that is, the particles have put themselves there in such a way that it works out, but you have not necessarily explained that they are there [28, p. 9101].

Gross did briefly return to the bootstrap approach with Veneziano’s discovery of the beta function formula, but quickly became disillusioned once again, this time by its inability to explain scaling. As a result, Gross quickly brought himself up to speed on quantum field theory (especially renormalization group techniques) to try to find an explanation of scaling within field theory. As we see in Chap. 9, he would return to a descendent of the bootstrap programme much later, in 1985, when he helped construct the heterotic string theory.

Though things obviously become near-exponentially complicated once we move outwards from the origins of the bootstrap approach and dual models, we can trace paths of several important string researchers from Mandelstam too, including Joseph Polchinski and Charles Thorn.

There were two quite distinct styles of physics associated with the West Coast (roughly: Berkeley, Caltech) and the East Coast (roughly: Chicago, Princeton, Harvard). In particular, the East Coast seems to have been less dominated by ‘physics gurus’ (if I might be permitted to use that term).Footnote 53 However, this is to ignore the European influence: there is clearly a strong European component, though this will really come to dominate the theory of strong interactions in the period around Veneziano’s presentation of his dual model.

This is, of course, very USA-centric, and much is missed. However, the influence spread across the Atlantic, especially to Cambridge University.Footnote 54 Mention should certainly be made too of the Japanese school. One of the initials of DHS duality (Richard Dolen) was based at Kyoto University for a time (at the Research Institute for Theoretical Physics). In his letters to Murray Gell-Mann (from 1966: in the Gell-Mann archives of Caltech [Box 6, Folder 20]) he explicitly mentions interactions with several local physicists that went on to do important work on dual models and string theory—including Keiji Kikkawa, who later visited Rochester in 1967.Footnote 55

Though it involves jumping ahead a little, much of the early detailed dual model work (including string models) took place at CERN. As has often been pointed out, this had much to do with the strong leadership and dual-model advocacy of Daniele Amati.Footnote 56 One could find David Olive (who would later take a post as a staff member, rather than a regular visitor, turning his back on a tenured position at Cambridge University), Peter Goddard, Ian Drummond, David Fairlie, and very many more centrally involved in the construction of string theory from the early dual resonance models.Footnote 57 Olive captures the hub-like dual model scene at CERN in the early 1970s as follows:

Amati had gathered together from around Europe a galaxy of young enthusiasts for this new subject as research fellows and visitors. This was possible as centres of activity had sprung up around Europe, in Copenhagen, Paris, Cambridge, Durham, Torino and elsewhere. I already knew Peter Goddard from Cambridge University who was in his second year as Fellow, Lars Brink from Chalmers in Gothenburg was just starting, as was Jöel Scherk from Orsay, in Paris, all as Fellows, and destined to be collaborators and, particularly, close friends. Also present as Fellows were Paolo Di Vecchia (who arrived in January 1972), Holger Nielsen, Paul Frampton, Eugène Cremmer, Claudio Rebbi and others. Many visitors came from Italy, Stefano Sciuto, Nando Gliozzi, Luca Caneschi and so on. Visiting from the United States for the academic year were Charles Thorn and Richard Brower. Summer visitors included John Schwarz, and later Pierre Ramond, Joel Shapiro, Korkut Bardakçi, Lou Clavelli and Stanley Mandelstam, all from the United States [42, p. 349].

The early phase involving dual models was a particularly interconnected one, then, and also one featuring very many collaborative efforts. Stefano Sciuti, who had earlier been a part of Sergio Fubini’s group in Turin, explicitly refers to the willingness to “join forces, cooperating rather than competing” as “fruit of the spirit of 1968” ([48], p. 216).

5 Summary

We have shown how the difficulties faced by quantum field theory in advancing beyond QED led to various models, one of which was Regge theory, with the addition of the dual resonance idea. This model achieved significant empirical successes, had several powerful theoretical virtues, and was therefore pursued with some excitement. We traced the story from Regge’s introduction of complex angular momentum into quantum mechanics, to its extension into the relativistic domain. This combined with ‘bootstrap’ physics according to which the properties of elementary particles, such as coupling constants, could be predicted from a few basic principles coupled with just a small amount of empirical input. This journey culminated in the finite energy sum rules of Dolen, Horn, and Schmid, which were elevated to the status of a duality principle. The primary researcher network guiding research in this period was fairly narrowly confined, and can be charted quite precisely, with Geoff Chew as a key hub leading an anti-QFT school, as far as strong interactions were concerned. The bulk of later developments which place Regge-resonance duality at the heart of hadron physics (and the true beginnings of string theory) take place across the Atlantic, at CERN. We turn to these in the next chapter in which we discuss the Veneziano (dual resonance) model and its many extensions and generalisations.