119
Views
0
CrossRef citations to date
0
Altmetric
Research Articles — Special Issue on the History of Fusion

Anthropic Importance of the “Bretscher State” in DT Fusion

ORCID Icon & ORCID Icon
Received 25 Jan 2024, Accepted 16 Mar 2024, Published online: 08 May 2024

Abstract

The term “Bretscher state” may not be as familiar as “Hoyle state,” but its anthropic importance cannot be overstated. In Big Bang nucleosynthesis, the deuterium-tritium (DT) fusion reaction 3H (d,n)4He, enhanced by the 3/2+ resonance due to the Bretscher state, is responsible for 99% of primordial 4He. While this fact has been known for decades, it has not been widely appreciated, and we recently proposed that its significance be commemorated by naming the 3/2+ state after Egon Bretscher, its discoverer. The importance of the resonant nature of the DT fusion reaction has been amplified by recent activities related to the production and use of terrestrial fusion including recent, net gain shots at the National Ignition Facility. Here, we aim to highlight the anthropic importance of the 4He-producing DT reaction that plays such a prominent role in models of nucleosynthetic processes occurring in the early universe. This primordial helium serves as a source for the subsequent creation of 25% of the carbon, 12C and other heavier elements that comprise a substantial fraction of the human body. Further studies are required to determine a better characterization of the amount of 12C than this lower limit of 25%. Some scenarios of core stellar nucleosynthetic yield of 12C suggest that even higher percentages of carbon from primordial helium are possible.

I. INTRODUCTION

We recently gave an account[Citation1,Citation2] of the earliest deuterium-tritium (DT) fusion discoveries, including the 1945–1946 Los Alamos Laboratory measurements of the DT cross section down to energies relevant to fusion energy applications (10 keV) by Egon Bretscher.[Citation3,Citation4] The 5He 3/2+ “Bretscher state” at 16.84 MeV, , leads to a hundredfold increase in the DT cross section. This resonant enhancement was a game-changer, opening up the potential for nuclear fusion technologies, as established by the recent inertial confinement fusion shot at the National Ignition Facility.[Citation5] It was in this context that we recently proposed, given the significance of the resonant enhancement in DTn+α [or, equivalently,  3H(d,n)4He], that the nuclear science community refer to 5He 3/2+ as the “Bretscher state.”[Citation6]

Fig. 1. Energy levels in the A = 5 system for the  3H(d,n)4He reaction. Resonant enhancement occurs when the energy of the deuterium d and tritium t reactants is close to the resonance energy, here 16.84 MeV for the 3/2+ Bretscher state. This happens to be within 50 keV (center-of-mass frame) of the dt separation energy (16.79 MeV), taking the resonance peak at Bretscher’s original value.[Citation3] (The current value of the resonance peak is a little higher at 65 keV.) Credit: S. Tasseff.

Fig. 1. Energy levels in the A = 5 system for the  3H(d,n)4He reaction. Resonant enhancement occurs when the energy of the deuterium d and tritium t reactants is close to the resonance energy, here 16.84 MeV for the 3/2+ Bretscher state. This happens to be within ∼50 keV (center-of-mass frame) of the dt separation energy (16.79 MeV), taking the resonance peak at Bretscher’s original value.[Citation3] (The current value of the resonance peak is a little higher at 65 keV.) Credit: S. Tasseff.

Here, we discuss the considerations of Ref.[Citation6] in greater detail including the anthropic implications of cosmological nucleosynthetic processes (see ) and the role of the resonant Bretscher state that contributed materially to the very existence of humans by its sourcing of the 4He fuel required for the manufacture of 12C. The present paper addresses events that occurred during Big Bang nucleosynthesis (BBN), the epoch covering the evolution of the early universe—from a few seconds to a few minutes after the Big Bang, some 13.8 billion years ago.

Fig. 2. A schematic of the dominant BBN pathways and the subsequent triple-alpha carbon formation in stars. The red curve shows the dominant path, which goes through the DT fusion reaction. Credit: S. Tasseff.

Fig. 2. A schematic of the dominant BBN pathways and the subsequent triple-alpha carbon formation in stars. The red curve shows the dominant path, which goes through the DT fusion reaction. Credit: S. Tasseff.

Work in the 1960s and 1970s by Wagoner et al.,[Citation7] Wagoner[Citation8,Citation9] (see ), and Peebles[Citation10] (extending earlier studies in Ref.[Citation11]) convincingly showed that—what has now become—the standard BBN model accounts for the primordial abundances of the light elements. This early model accounts fairly accurately for the light isotope abundances with roughly 75% hydrogen (1H), 25% helium (4He) by mass,[Citation11] and trace amounts of other nuclides, all created within the first few minutes after the Big Bang by the processes depicted in . A more recent study at California Institute of Technology (Caltech) by Smith et al.[Citation12] provided a detailed sensitivity study to determine the most important nuclear reaction pathways explored in the epoch of the Big Bang, during the first few hundred seconds, using accurate representations of thermonuclear cross sections.

Fig. 3. Robert V. Wagoner, William A. Fowler, Fred Hoyle, and Donald D. Clayton in February 1967, between Sloan Laboratory and West Bridge Laboratory on the Caltech campus.

Fig. 3. Robert V. Wagoner, William A. Fowler, Fred Hoyle, and Donald D. Clayton in February 1967, between Sloan Laboratory and West Bridge Laboratory on the Caltech campus.

The present account is based on Smith et al.’s calculations, which are supported by more recent, detailed studies such as those of Grohs et al.[Citation13] and Pitrou et al.[Citation14] In order to commence our discussion of the production of 4He in BBN, the dominance of the DT reaction, and the importance of the Bretscher state, we first review the nuclear reaction history of the early universe as a function of its temperature and highlighting the role of entropy in its evolution.

II. NUCLEAR REACTION HISTORY OF THE EARLY UNIVERSE

The abundance history of the light isotopes in the early universe,Footnotea determined numerically, is shown in , along with the entropy per baryon. The derivative with respect to time is given as

(1) dstotdt=Tcm32π2nbi=160ϵ2Cνi[fj]  ×[ϵTcmT+ln(fi1fi)],(1)

Fig. 4. The light isotope abundances Y=YA,Z [see EquationEq. (3)] (upper panel: dimensionless, measured relative nucleon number density) and the entropy per baryon stot (lower panel: dimensionless) as a function of the temperature of the early universe covering a range of times measured from the Big Bang from 1 to 100 s.

Fig. 4. The light isotope abundances Y=YA,Z [see EquationEq. (3(3) YA,Z=gAA3/2222π7/2gs45TcmmN3/2A−1YpZYnA−Z×1stotA−1eBA,Z/Tcm,(3) )] (upper panel: dimensionless, measured relative nucleon number density) and the entropy per baryon stot (lower panel: dimensionless) as a function of the temperature of the early universe covering a range of times measured from the Big Bang from ∼1 to ∼100 s.

where nb = baryon number density per unit volume; fi(ϵ,t) = neutrino i=1,2,3 and antineutrino i=4,5,6 momentum probability distribution functions; Cνi = rate of collisions of neutrinos and antineutrinos, which are computed in the solution of the neutrino-antineutrino Boltzmann equations. The entropy per baryon is expressed as a function of the (comoving) temperature Tcm of the early universe and T, the temperature of charged leptons and the baryons (n and p).

The temperatures involved are very large, even by solar scales, and this necessarily implies that a very large entropy is carried by the elementary constituents of the early universe. The entropy, as a measure of the disorder of the system, is an extensive quantity, depending on the number of particles, and is therefore dominated by the photons and the neutrinos. (The nucleons collectively do not carry much of the entropy since there is only 1 nucleon for every 1010 photons or neutrinos.) The energy units on the x-axis of are proportional to temperature (via the Boltzmann constant kB1=11.6 GK/MeV, where GK means 109 kelvin). Reading the figure from left to right, the temperature decreases from a time when the age of the universe was about 1 s old with temperature Tcm(\lsim 1s)\gsim10MeV100 GK to a time when the temperature is too low for the positively charged ions to overcome the Coulomb repulsion. Here, at Tcm(\gsim 100s)\lsim 0.01MeV0.1 GK, the nuclear reactions have all but ceased.

Returning to the leftmost, higher-temperature part , we consider the initial conditions that feed the nuclear reactions that bind neutrons and protons in nuclei. The central points are that the final, primordial abundances are limited by the number of available neutrons and that nearly all the neutrons end up being bound in 4He nuclei (see ). The first point is a consequence of the nature of the weak interactions that interconvert np[Citation15] and the fact that neutron-destroying reactions dominate. This is mainly because the neutron mass is about 1.3 MeV larger than that of the proton, independent of the fact that it is unstable. (The neutron does not have time to decay much before the nuclear reactions begin at a few seconds after the Big Bang.) The ratio of (the number density of) neutrons to protons is governed by the weak interactions of nucleons with charged (electrons and positrons) and neutral (neutrinos and antineutrinos) leptons.

TABLE I Final Primoridal Abundances (Dimensionless) Relative to Nucleon Mass Density (1H and 4He) or Relative to Nucleon Number Density (2H, 3H, 3He, 6Li, 7Li, and 7Be)

The assumption of chemical equilibrium between the neutrons n and protons p gives an approximate expression for the ratio of their number densities: n/pexp(δmnp/Tcm), where δmnp1.293 MeV is the mass of the neutron less the mass of the proton. At high temperature, Tcmδmnp, the baryonic matter comprises equal parts neutrons and protons n/p1. At a lower temperature of about Tcm\lsim0.9 MeV, i.e., the “weak freeze-out” temperature, the reactions that interconvert np cease with the n/p fraction at weak freeze-out 1/6.

The nuclear abundances (measured relative to the nucleon or baryon number density) are then set—this is the second, central point made earlier—by this initial n/p1/6 condition and a phase in the evolution called nuclear statistical equilibrium (NSE). The NSE obtains when all the forward (say, np) and reverse (γdnp) nuclear rates are very nearly equal and very large compared to the rate of expansion of the universe. The rate of expansion is set by the Hubble parameter H and can reach upward of 100 Hz (in inverse seconds) near Tcm  10 MeV.

Under these extreme conditions of high temperature, the system is held in a quasi-equilibrium state by the extremely high rates of the nuclear reactions, and a condition close to detail balance holds. That is, for each nuclear reaction that fuses neutrons and protons (nucleons) into heavier nuclei with nucleon numbers A>2, another reaction immediately follows that breaks up the heavier A>2 nucleus into its component nucleons. In these near-equilibrium conditions, the Gibbs free energy G=aμaNa is conserved by the nuclear reactions. Here, μa and Na are the chemical potentials and numbers of species a, respectively, where a indexes the individual nuclides (a=n,p,d, 3H,3He,4He,). These conditions (of chemical equilibrium) obtain at high temperatures while the entropy density (lower panel, ) is nearly constant for Tcm\gsim1 MeV and are expressed in the form

(2) μA,Z=(AZ)μn+Zμp,(2)

where A,Z = atomic weight and number of protons Z of the nuclide A; μA,Z = chemical potential of the nuclide (nucleus) with Z protons and AZ neutrons (for example,  4He has A=4,Z=2,N=2); μn = neutron chemical potential; μp = proton chemical potential. EquationEquation (2) can be used to determine the element abundances, shown in as dashed curves, for high temperatures relevant for NSE. For Tcm\lsim1 MeV, the system departs from the quasi-equilibrated NSE state as the nuclear reaction rates drop below the Hubble rate H.

Fig. 5. The light isotope abundance Y, relative to the nucleon number density, calculated numerically (solid curves) compared to the NSE values (dashed curves) for isotopes with charge 4, shown with the entropy per baryon (lower panel), as in . The dashed and solid curves overlap when NSE obtains.

Fig. 5. The light isotope abundance Y, relative to the nucleon number density, calculated numerically (solid curves) compared to the NSE values (dashed curves) for isotopes with charge ≤4, shown with the entropy per baryon (lower panel), as in Fig. 4. The dashed and solid curves overlap when NSE obtains.

Before discussing the details of the nuclear reactions that give rise to the abundances in and , we point out some general features of the abundance history graph.Footnoteb

During the phase of NSE, EquationEq. (2)Footnotec may be solved to give

(3) YA,Z=gAA3/2222π7/2gs45TcmmN3/2A1YpZYnAZ×1stotA1eBA,Z/Tcm,(3)

where the abundance YA,Z is the ratio of the number density of nuclide (A,Z) relative to the nucleon (or baryon) number density. The quantities gA=2JA+1, where JA is the spin of nuclide of mass A and gs=10.75 count the number of effective spin degrees of freedom of the nuclide of mass A and the plasma constituents. (See Ref.[Citation18], p. 87ff, for a detailed discussion.) Suffice it to say here that this factor is only weakly dependent on both the properties (A,Z) of the nuclei and the temperature Tcm. The factor in the second line of EquationEq. (3), stot1Aexp(BA,Z/Tcm), is however very strongly dependent on the temperature Tcm and the atomic weight A. It represents a competition between the energetics of nuclear binding and the disordering effect of the high-entropy environment created by the huge number of photons at high temperature that we will call the photon bath.

The strong dependence on Tcm is owed to the exponential factor that features the nuclear binding energy (see ) BA,Z=Zmp+(AZ)mnmA,Z, where mp,mn,andmA,Z are the mass of the proton, neutron, and nucleus with atomic weight A and charge Z. The strong dependence on A is owed to the fact that the entropy per baryon stot is a very large (dimensionless) number.Footnoted The entropy in the early universe, dominated by contributions from the electroweak components (photons, charged leptons, and neutrinos), is of order stot1010, as seen in the lower panel of . The origin of the exponential factor, which favors nuclides with larger total binding energies, is a consequence of the Boltzmann factor exp(E/T), which arises in the (canonical) thermal average in the determination of the nuclide density. The entropic effect, which suppresses nuclides with higher numbers A of nucleons, is due to the influence of the photon bath at high temperature (\gsim1 MeV) and the tendency for this entropy to be transferred to the forming nuclear system thereby reducing the likelihood that larger A systems will form.

TABLE II Binding Energy of Some Light Isotopes*

These tendencies are clearly visible in the plot of the nuclear abundances in both and . At temperatures Tcm far above 1 MeV, the entropic effect suppresses each nuclide in NSE by powers of A: Y 4He<<Y 3H,3He<<Y 2H. The rate of increase of the abundances near Tcm\lsim1 MeV is governed by the value of BA,Z with the rate of 4He production largest.

Moving to the right-hand side of the figure, at temperatures Tcm\lsim 30 keV, the abundances are essentially constant except for that of the neutron, which falls off due to free-neutron decay with a half-life of about 10 min.Footnotee This is a consequence of the complete freeze-out of all of the electroweak and nuclear reactions, a result of the Hubble expansion cooling the reactive plasma, which transmute nuclides into one another. The final, unchanging values of the nuclear abundances determine what is referred to as the primordial abundances. The most abundant primordial nuclide (with A>1) is 4He, as expected from the foregoing discussion of the role of its large binding energy, making up about one-fourth of the baryonic matter by mass. Most of the rest of the mass fraction is accounted for by hydrogen.

The remaining trace nuclides are essentially the ash left in the thermonuclear furnace of the early universe. The primordial deuterium abundance, with number fraction Y 2H (of baryonic matter), is about 1 part in 105, and the mass 3 nuclei are down about one to several orders of magnitude below 2H. The 7Li nuclide,Footnotef infamous because of a persistent discrepancy between the abundance predicted in the standard picture (given here) and observations that find a factor value of between two to five times smaller, is about 1 in 1010. This discrepancy indicates either a problem with the observations or a physical mechanism that destroys the mass 7 nuclides (7Li and 7Be) but has only marginal effect on the other well-measured nuclides [2H and 4He]; see Ref.[Citation20] for a recent discussion.

In the intermediate-temperature region, from 1 MeV \gsimTcm\gsim 30 keV, starting at the higher-temperature end of this region, the evolution is determined by a sequence of partial nuclear reaction freeze-outs, which are detectable by the change of the curvature of the abundance curves from positive to negative and back to positive. The first of these regions occurs for the 4He abundance at about Tcm0.6 MeV. It is at this point that the NSE abundance (dashed curve in ) diverges from the numerical simulation and corresponds to a point in the evolutionary history of the early universe when the weak interactions that kept the neutrons and protons close to equilibrium become slow enough that the NSE no longer obtains. At this point, the 4He abundance drops off its NSE curve and begins to more closely track the slower mass 3 nuclides (3H and 3He), which are the source for reactions that produce the 4He. Subsequent curvature transitions are visible in the numerical simulation and correspond to similar rate freeze-out effects.

We now turn to the individual production and destruction mechanisms responsible for the major isotopes produced in Big Bang (or primordial) nucleosynthesis.

III. INDIVIDUAL PRODUCTION AND DESTRUCTION MECHANISMS

III.A. BBN Deuterium Production Via p(n,γ)2H

At the early stages of the Big Bang, as previously mentioned, the temperature was sufficiently high for neutron and proton number densities to be equal since np transmutations, mediated by the weak interaction, maintained their equilibrium. Owing to the high-entropy, energetic photon bath environment of the early universe (photons outnumber nucleons by about 1010 to 1), the electromagnetic reaction n(p,γ)2H and its reverse  2H(γ,n)p remain in equilibrium for temperatures down to about Tcm70 keV, far below that expected simply on the basis of the binding factor expB2,1/Tcm in EquationEq. (3). This is visible in when the dashed, blue NSE curve for 2H diverges from the solid, blue numerical curve. At this stage, the free neutron supply (see of Ref.[Citation12]) is being rapidly depleted by other, highly efficient, reactions that ultimately pack the neutrons predominantly into 4He nuclei, and the deuteron 2H production begins to fall below its NSE track. The 2H destructive reactions, such as  2H(γ,n)p,  2H(d,n)3He,  2H(d,p)3H, and perhaps most importantly  3H(d,n)4He, ablate some of the previously produced deuterium, reducing its abundance (or mass fraction) by two or three orders of magnitude compared to its peak value near Tcm  60 keV.Footnoteg

III.B. BBN Tritium and 3He Production Via DD

The NSE curves in for 3H and 3He are coincident with their numerically determined values until Tcm  0.2 MeV where the mass 3 destructive reactions begin to overcome those that produce mass 3. The more tightly bound 3H (compared to3He; see ) gives it an advantage during NSE, as seen in the exponential factor in EquationEq. (3). The two branches of the mass 3 producing deuterium-deuterium (DD) fusion reactions, i.e.,  2H(d,p)3H and  2H(d,n)3He, shown in , produce neutrons, protons, and the mass 3 nuclides. These fusion reactions occurred at nearly equal rates (a consequence of isospin symmetry) producing tritium (T or 3H) and 3He in nearly equal amounts. Again, 3H has an advantage over the latter, 3He nuclide, since it is more readily transmuted by the neutron-induced, charge neutral  3He(n,p)3H reaction than its reverse  3H(p,n)3He, which is Coulomb suppressed. These same DD reactions were first observed in the laboratory in 1934 by Oliphant, Harteck, and Rutherford, and the cross sections were measured at the Metallurgical Laboratory in Chicago (1942–1944) and at Los Alamos Laboratory by Bretscher.[Citation1,Citation2]

III.C. BBN 4He Production Via DT

The 4He mass fraction in (black curves), previously discussed in terms of NSE and the partial freeze-out mechanism, is seen to fall out of equilibrium with the mass 3 nuclides near Tcm150 keV and subsequently to enjoy an explosive acceleration in its production near Tcm65 keV. This distinctive and perhaps surprising feature occurs at precisely the same energy where the temperature-dependent reactivity σv reaches its maximum; see in the companion paper in this special issue.[Citation2] This accelerated production is due to resonance enhancement of 4He production by the 3/2+ Bretscher state in the 5He compound system, which has a corresponding peak cross section of 5 b at 65 keV. We might pause to emphasize that the large magnitude of the cross section, compared to other light nuclear reactions, is unparalleled and clearly the dominant effect in the production of 4He.

Fig. 6. Egon Bretscher (foreground, left) and Robert Serber (foreground, right) at the 1946 Nuclear Physics Conference held in Los Alamos.

Fig. 6. Egon Bretscher (foreground, left) and Robert Serber (foreground, right) at the 1946 Nuclear Physics Conference held in Los Alamos.

In fact, the calculations of Smith et al. (see in Ref.[Citation12]) show that 99% of the 4He was produced via  3H(d,n)4He, labeled “DT fusion” in . The remaining 1% of primordial 4He came from the  2H(3He,p)4He reaction, which benefits from the same mirror 3/2+ resonance but is suppressed by the larger Coulomb repulsion between D and 3He. As we pointed out previously in Sec. III.B. on mass 3 production, much of this 3He was rapidly converted to tritium by the (n,p) reaction, which becomes another source for 4He via the very fast DT fusion reaction.[Citation20] The importance of DT in BBN was noted by Hupin et al.[Citation21] in their interesting paper on ab initio cross-section calculations. A central point of our current considerations is that 4He, the fuel for further nucleosynthesis of heavier elements in stars, which constitutes about 25% by mass of the baryonic matter in the universe, was dominantly made via resonant DT fusion in BBN.

Fig. 7. DT and DD cross sections in the incident deuteron projectile laboratory frame, from EDA R-matrix code analyses, including a thought experiment DT calculation in which the A = 5 3/2+ Bretscher state is removed.

Fig. 7. DT and DD cross sections in the incident deuteron projectile laboratory frame, from EDA R-matrix code analyses, including a thought experiment DT calculation in which the A = 5 3/2+ Bretscher state is removed.

III.D. Stellar Nucleosynthesis of 4He

In his Nobel Prize–awarded work, “Energy Production in Stars,”[Citation22] Bethe laid out that the only two reaction chains, consistent with known nuclear cross sections, that convert four protons and two electrons to a helium nucleus are the proton-proton (p-p) chain and the carbon-nitrogen-oxygen (CNO) cycle. The former is important in less massive stars (below 1.2 M), and the latter is important in heavier ones (above 1.2 M). Both have the effect of transmuting hydrogen to helium without changing the chemical composition of other, heavier, elements. In the CNO cycle, some 12C must be present to seed the reaction chain, and the 12C nucleus is reproduced at the end of the cycle by the reaction  15N(p,α)12C.

Our current, anthropic considerations lead us to ask the following question: In an idealized scenario where stars are formed only from primordial material (75% hydrogen and 25% helium), what proportion of the resulting helium produced by whatever stellar nucleosynthetic mechanism comes from BBN or stellar processing? The then ensuing process of igniting, in sufficiently massive stars, the 4He in triple-alpha burning reaction takes place resulting in the production of carbon, the foundation of all organic chemical compounds on our planet. Such sufficiently massive stars are twice as large as the sun as they evolve into red giants, when their cores contract and heat further, to about six times that of the sun.[Citation23,Citation24]

Similar to the DT reaction that produces 4He, the triple-alpha reaction is resonantly enhanced. In the triple-alpha process, two alpha particles form the short-lived (1016 s) 8Be compound nucleus. Should a third alpha particle find the ephemeral 8Be before it decays, the reaction  8Be(α,2γ)12C occurs with branching Γrad/Γ  4.1(1)×104 dominated by the 12C resonance with excitation energy 7.65 MeV and spin parity 0+, the 12C Hoyle state.[Citation25] Then, more 4He was catalyzed by the CNO cycle, and heavier element nucleosynthesis could ensue through progressive captures of α particles on the chain of nuclei starting with 12C. Although the CNO stellar nucleosynthesis reactions increased the 4He content in the universe, the increase is smaller, compared with the initial BBN amount,[Citation18] even in the case of complete hydrogen burning to helium.

It is interesting to ask what percentage of the 4He that is converted to carbon and heavier elements came from BBN-produced helium compared to stellar nucleosynthesis–produced helium, where the hydrogen was transmuted to helium through p-p chain and/or CNO processes. Complete hydrogen burning in stellar nucleosynthesis places a rigorous lower bound of 25% 4He from BBN, required by the fact that the number of neutrons and protons is conserved. This important point bears reiteration: If the primordial material from which carbon is synthesized had all of its hydrogen converted to helium, then one would have the lower limit, 25% BBN helium. Otherwise, with incomplete burning of the primordial hydrogen—which is likely—the percentage of 4He would be higher. A more precise estimate requires future study via detailed simulations.[Citation23]

Many features of the universe we inhabit were therefore produced through the resonance-enhanced DT fusion reaction, notably 25% of the universe’s mass being He. Also, at least 25% of the mass in elements comprising 90% of our body’s mass (the nonhydrogen component) was created in BBN DT fusion through the Bretscher state—a substantial fraction!

III.E. The Bretscher State

Because the 5He 3/2+ 16.84-MeV state discovered by Bretscher[Citation3,Citation4] is responsible for the fast rate of the DT reaction and the concomitant implications described in this paper, we propose that the nuclear fusion community refer to it as the Bretscher state, in analogy with the 12C Hoyle state. The magnitude and importance of the DT resonant enhancement effect is evident in , where we have used an R-matrix reaction analysis code (the Los Alamos National Laboratory EDA code[Citation26]) to calculate the DT cross section with and without the 3/2+ Bretscher state. If there was no such state, DT would be similar to the DD cross section.

IV. CONCLUSION

Given the obscurity of the role of the Bretscher state and its singular anthropic importance, in closing, we should briefly revisit the history of its discovery by Bretscher and collaborators. The Los Alamos National Laboratory archives, housed in the National Security Research Center, hold a variety of original materials available nowhere else, including monthly progress reports from F-Division, the division led by Enrico Fermi during the Manhattan Project. These reports record that Bretscher and French obtained their first DT data (with tritons impinging on a heavy ice target) in July 1945 (for a center-of-mass energy of 16 keV). A detailed characterization of the DT resonance was given in Bretscher’s Los Alamos Laboratory report LA-582, dated November 15, 1946.[Citation3] An image, extracted from this report, is given in , and one can see that this document is identical to the 1949 version published later in Physical Review.[Citation4] (This 1946 Los Alamos Laboratory report was initially classified as a secret document; its open publication was delayed until 1949.) The report establishes the resonant nature of the 2H(t,n)4 (T on D) reaction and provides two approaches to determine the resonance energy. One (see ) gave the resonance energy in terms of the incident triton energy Et124 keV. The other method gave it as Et343 keV. In the center of mass, these values are equivalent to 50 and 137 keV, respectively, spanning the known value today of 65 keV. Although Bretscher was the first to document and quantify the resonance’s energy, we think it very likely that Bethe would have recognized the likely presence of a resonance as early as 1943 when the first DT cross-section data were obtained at Purdue University[Citation1,Citation2] (though at energies above the resonance’s peak). The spin of the resonance was correctly identified by 1952; see our discussion on the evolving understanding of the 3/2+ resonance in Ref.[Citation2].

Fig. 8. An extract from Bretscher and French’s paper LA-582,[Citation3] November 15, 1946, providing the first identification of the DT resonance.

Fig. 8. An extract from Bretscher and French’s paper LA-582,[Citation3] November 15, 1946, providing the first identification of the DT resonance.

While Bretscher is generally credited with correctly anticipating the properties of 239Pu,[Citation2,Citation27] his discovery of the DT resonance has gone largely unappreciated. His 1945–1946 DT measurements, which were published in Physical Review in 1949,[Citation4] have been cited only 54 times, compared with the 460 citations of Hoyle’s paper. This is—perhaps strangely—similar to the fact that the first-ever observation of DT fusion, in 1938 by Ruhlig,[Citation28] had never been cited until our recent paper.[Citation1] It is within this context that we propose the naming of the 3/2+ DT resonance as the Bretscher state, in the spirit of its being anthropically important, like the Hoyle state.

In Ref. [Citation2], we present a brief biography of Bretscher, which includes his early work at Cambridge; his time at Los Alamos National Laboratory; his friendship with Oliphant, Bloch, and Staub; and his later leadership roles in nuclear energy research at Harwell.

Acknowledgments

We acknowledge useful discussions with Michael Smith, Jonathan Katz, Joyce Guzik, Michael Bernardin, Evan Grohs, Gerry Hale, George Fuller, Craig Carmer, and Tom Kunkle.

Disclosure Statement

No potential conflict of interest was reported by the author(s).

Additional information

Funding

This work was supported by the U.S. Department of Energy through the Office of Science [13313917] and Los Alamos National Laboratory. Los Alamos National Laboratory is operated by Triad National Security, LLC, for the National Nuclear Security Administration of U.S. Department of Energy (Contract No. 89233218CNA000001).

Notes

a The abundance history is determined from a system of coupled nonlinear ordinary differential equations in time that account for the general-relativistic expansion of the universe, its thermodynamics, and the strong and electroweak nuclear reaction rates.[Citation8,Citation9]

b We thank George Fuller for emphasizing to us the importance of the role of entropy in the state of NSE.[Citation16,Citation17]

c In the context of atomic and plasma physics, this relation is known as the Saha equation and allows one to compute the ionization state of the gas as a function of temperature in terms of the energy levels of the atom.

d We might qualify use of the phrase “a very large number” as the black hole entropy, given by the Bekenstein-Hawking relation, has the much higher value of ~1077 for a one solar mass object.

e The time after the Big Bang is inversely proportional to the square of the temperature: tTcm2.

f We sum the abundances of the mass 7 nuclides, 7Li and 7Be, since 7Be decays with half-life ~53 days due to  7Be(e,νe)7Li, in order to obtain the 7Li primordial abundance.

g Perhaps it is worth noting that BBN production of deuterium differs from its production in stars where there are no free neutrons. Stellar nucleosynthesis of deuterium relies on the relatively much slower weak interaction p+p2H+e++νe.

References

  • M. B. CHADWICK et al., “The Earliest DT Fusion Discoveries, 1938–1952,” LA-UR-22-32324, WRL-100, Los Alamos National Laboratory (2023); https://arxiv.org/abs/2302.04206. Nuclear News, 66, 70 (Apr. 2023).
  • M. CHADWICK et al., “Early Nuclear Fusion Cross Section Advances 1934–1952 & Comparison to Today’s ENDF Data,” Fusion Sci. Technol., X, x (2024). LA-UR-22-23373 Version 4, Los Alamos National Laboratory ( 2022).
  • E. BRETSCHER and A. P. FRENCH, “Low Energy Cross Section of the D-T Reaction and Angular Distribution of the Alpha Particles Emitted,” LA-582, Los Alamos Laboratory (1946).
  • E. BRETSCHER and A. P. FRENCH, “Low Energy Cross Section of the D−T Reaction and Angular Distribution of the Alpha-Particles Emitted,” Phys. Rev., 75, 1154 (1949); https://doi.org/10.1103/PhysRev.75.1154.
  • H. ABU-SHAWAREB et al., “Lawson Criteria for Ignition Exceeded in an Inertial Fusion Experiment,” Phys. Rev. Lett., 129, 7, 075001 (2022); https://doi.org/10.1103/PhysRevLett.129.075001.
  • M. W. PARIS and M. B. CHADWICK, “Big Bang Fusion 13.8 Billion Years Ago and Its Importance Today,” Nuclear News, 66, p. 116 (Aug. 2023).
  • R. V. WAGONER, W. A. FOWLER, and F. HOYLE, “On the Synthesis of Elements at Very High Temperatures,” Astrophys. J., 148, 3 (1967); https://doi.org/10.1086/149126.
  • R. V. WAGONER, “Synthesis of the Elements Within Objects Exploding from Very High Temperatures,” Astrophys. J. Suppl., 18, 247 (1969); https://doi.org/10.1086/190191.
  • R. V. WAGONER, “Big-Bang Nucleosynthesis Revisited,” Astrophys. J., 179, 343 (1973); https://doi.org/10.1086/151873.
  • P. J. PEEBLES, “Primeval Helium Abundance and the Primeval Fireball,” Phys. Rev. Lett., 16, 410 (1966); https://doi.org/10.1103/PhysRevLett.16.410.
  • R. ALPHER, H. BETHE, and G. GAMOW, “The Origin of Chemical Elements,” Phys. Rev., 73, 803 (1948); https://doi.org/10.1103/PhysRev.73.803.
  • M. S. SMITH, L. H. KAWANO, and R. A. MALANEY, “Experimental, Computational, and Observational Analysis of Primordial Nucleosynthesis,” Astrophys. J. Suppl., 85, 219 (1993); https://doi.org/10.1086/191763.
  • E. GROHS et al., “Neutrino Energy Transport in Weak Decoupling and Big Bang Nucleosynthesis,” Phys. Rev. D, 93, 8, 083522 (2016); https://doi.org/10.1103/PhysRevD.93.083522.
  • C. PITROU et al., “Precision Big Bang Nucleosynthesis with Improved Helium-4 Predictions,” Phys. Rep., 754, 1 (2018); https://doi.org/10.1016/j.physrep.2018.04.005.
  • E. GROHS and G. M. FULLER, “The Surprising Influence of Late Charged Current Weak Interactions on Big Bang Nucleosynthesis,” Nucl. Phys. B, 911, 955 (2016); https://doi.org/10.1016/j.nuclphysb.2016.08.034.
  • G. M. FULLER, “Quantum Effects on Precision Cosmological Observations: Introduction & Overview,” presented at QuEPCO 2017, Santa Fe, New Mexico, August 22–24, 2017.
  • M. S. TURNER, “Understanding BBN: The Physics and Its History” (2022); https://arxiv.org/abs/2111.14254.
  • E. W. KOLB and M. S. TURNER, The Early Universe, Vol. 69 (1990); https://doi.org/10.1201/9780429492860.
  • M. WANG et al., “The AME 2020 Atomic Mass Evaluation (II). Tables, Graphs and References,” Chin. Phys. C, 45, 3, 030003 (2021); https://doi.org/10.1088/1674-1137/abddaf.
  • B. D. FIELDS, P. MOLARO, and S. SARKAR, “Review of Particle Physics: Big Bang Nucleosynthesis,” PTEP, 2020, 8, 083C01 (2020) ( and 2021 update); https://doi.org/10.1093/ptep/ptaa104.
  • G. HUPIN, S. QUAGLIONI, and P. NAVRATIL, “Ab Initio Predictions for Polarized Deuterium-Tritium Thermonuclear Fusion,” Nat. Commun., 10, 1, 351 (2019); https://doi.org/10.1038/s41467-018-08052-6.
  • H. A. BETHE, “Energy Production in Stars,” Phys. Rev., 55, 5, 434 (1939); https://doi.org/10.1103/PhysRev.55.434.
  • J. GUZIK, Personal Communication (2023).
  • E. M. BURBIDGE et al., “Synthesis of the Elements in the Stars,” Rev. Mod. Phys., 29, 4, 547 (1957); https://doi.org/10.1103/RevModPhys.29.547.
  • F. AJZENBERG-SELOVE, “Energy Levels of Light Nuclei A = 11–12,” Nucl. Phys. A, 506, 1, 1 (1990); https://doi.org/10.1016/0375-9474(90)90271-M.
  • G. HALE and M. PARIS, “Data Covariances from R-Matrix Analyses of Light Nuclei,” Nucl. Data Sheets, 123, 165 (2015); https://doi.org/10.1016/j.nds.2014.12.029.
  • M. B. CHADWICK, “Nuclear Science for the Manhattan Project and Comparison to Today’s ENDF Data,” Nucl. Technol., 207, S1, S24 (2021); https://doi.org/10.1080/00295450.2021.1901002.
  • A. J. RUHLIG, “Search for Gamma-Rays from the Deuteron-Deuteron Reaction,” Phys. Rev., 54, 4, 308 (1938); https://doi.org/10.1103/PhysRev.54.308.