12.07.2015 Views

Download Thesis in Pdf Format - Theoretical Nuclear Physics and ...

Download Thesis in Pdf Format - Theoretical Nuclear Physics and ...

Download Thesis in Pdf Format - Theoretical Nuclear Physics and ...

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

FACULTEIT WETENSCHAPPENA RELATIVISTIC EIKONALDESCRIPTION OF NUCLEONPROPAGATION THROUGH NUCLEIBart Van Overmeirereal part of S IFSI0.80.60.40.2150θ ( o )1005012345r (fm)60Promotor: Prof. dr. Jan RyckebuschProefschrift <strong>in</strong>gediend tot het behalen van de academische graad vanDoctor <strong>in</strong> de Wetenschappen: NatuurkundeUniversiteit GentFaculteit WetenschappenVakgroep Subatomaire en Stral<strong>in</strong>gsfysicaAcademiejaar 2006-2007


Voor Barbara en Jules


Contentsiv


Chapter 1IntroductionEver s<strong>in</strong>ce the discovery of the atomic nucleus by Rutherford <strong>in</strong> the early 20th century, nuclei<strong>and</strong> their constituent particles have been the subject of a great deal of scientific research. The<strong>in</strong>terest <strong>in</strong> nuclear physics stems from the fact that the nucleus is a unique form of matter,consist<strong>in</strong>g of many protons <strong>and</strong> neutrons <strong>in</strong> close proximity. As such, the nucleus providesa microscopic laboratory to study different forces of nature: the strong, electromagnetic, <strong>and</strong>weak <strong>in</strong>teraction. Moreover, nuclear matter accounts for more than 99.9% of the mass <strong>in</strong> thevisible universe. Consequently, nuclear physics plays a crucial role <strong>in</strong> the underst<strong>and</strong><strong>in</strong>g of ourworld, from the <strong>in</strong>f<strong>in</strong>itesimal to the astronomical. F<strong>in</strong>ally, there are many practical applicationsof nuclear physics, <strong>in</strong> energy production, medical diagnosis <strong>and</strong> treatments, to name a few.A powerful tool to extract <strong>in</strong>formation about nuclear structure <strong>and</strong> the properties of boundnucleons is the scatter<strong>in</strong>g of leptonic or hadronic probes from nuclei. The advantages of us<strong>in</strong>gleptons as probes arise from the fact that they <strong>in</strong>teract weakly with the nucleus. This enables theleptons to probe the entire nuclear volume without significantly disturb<strong>in</strong>g the target object. Italso means that the electroweak <strong>in</strong>teraction can be treated perturbatively. Leptonic scatter<strong>in</strong>g,however, has the disadvantage that the cross sections are much smaller than those for purelyhadronic reactions. Thus, hadronic scatter<strong>in</strong>g has a large “discovery” potential, whereas leptonicreactions are used for precision experiments. Hadronic scatter<strong>in</strong>g is more challeng<strong>in</strong>gtheoretically because of the strong <strong>in</strong>teraction with nuclear matter. Moreover, hadronic probesmostly sample the nuclear surface.Quasielastic Scatter<strong>in</strong>g <strong>and</strong> the A(p, pN) ReactionThe ma<strong>in</strong> focus of this work will be on exclusive proton-nucleus scatter<strong>in</strong>g <strong>in</strong> the quasielasticregion. In the scatter<strong>in</strong>g spectrum of a hadronic probe from the nucleus, the quasielastic regionappears as a wide peak at ω ≈ Q 2 /2M N , with (ω, ⃗q) the energy <strong>and</strong> momentum transfer, Q 2 =|⃗q| 2 − ω 2 , <strong>and</strong> M N the nucleon mass. One expects that <strong>in</strong> this region, the probe <strong>in</strong>duces theknockout of a s<strong>in</strong>gle nucleon. In exclusive measurements, the ejectile <strong>and</strong> the knocked-outnucleon are detected <strong>in</strong> co<strong>in</strong>cidence <strong>and</strong>, hence, more detailed <strong>in</strong>formation about the reaction1


2mechanism is obta<strong>in</strong>ed than <strong>in</strong> <strong>in</strong>clusive experiments.To first order, quasielastic scatter<strong>in</strong>g can be understood as scatter<strong>in</strong>g from the bound nucleons,i.e., the <strong>in</strong>teraction with the target nucleus is assumed to occur through the <strong>in</strong>dividualnucleons. This is the so-called impulse approximation (IA). Hence, a quasielastic A(p, pN) reactioncan be described as follows: The energy ω <strong>and</strong> momentum ⃗q transferred by the imp<strong>in</strong>g<strong>in</strong>gproton to the nucleus are predom<strong>in</strong>antly absorbed by a s<strong>in</strong>gle bound nucleon, which is subsequentlyknocked out of the target nucleus. Thus, the struck nucleon is assumed to co<strong>in</strong>cidewith the ejected one. S<strong>in</strong>ce the ejectile absorbs the major fraction of the transferred energy, theresidual A − 1 nucleus is left <strong>in</strong> its ground state or <strong>in</strong> a low-ly<strong>in</strong>g excited state.Over the last five decades, quasielastic nucleon-knockout reactions have been extensively<strong>in</strong>vestigated. For a review of the research <strong>in</strong>to proton-<strong>in</strong>duced s<strong>in</strong>gle-nucleon emission offnuclear targets, the reader is referred to Refs. [1–3]. The basic idea beh<strong>in</strong>d the A(p, pN) modelsis the IA. This relatively simple reaction mechanism of one “hard” nucleon-nucleon collisionis, however, obscured by the “soft” <strong>in</strong>itial- <strong>and</strong> f<strong>in</strong>al-state <strong>in</strong>teractions (IFSI) of the <strong>in</strong>cident <strong>and</strong>two outgo<strong>in</strong>g nucleons with the nuclear medium. A sketch of the A(p, 2p) reaction is given <strong>in</strong>Fig. 1.1. Any model for A(p, 2p) <strong>and</strong> A(p, pn) reactions needs to address at least three issues:• The hard scatter<strong>in</strong>g that leads to the ejection of the struck nucleon.• The structure of the target <strong>and</strong> residual nucleus.• The distort<strong>in</strong>g mechanisms of the soft small-angle IFSI.K 1ppP 1A-1FSIpISIK 2AK A K A-1Figure 1.1 Schematic representation of the A(p, 2p) reaction. The <strong>in</strong>com<strong>in</strong>g proton undergoes “soft”<strong>in</strong>itial-state <strong>in</strong>teractions with the target before knock<strong>in</strong>g out a bound proton through the primary highmomentum-transferpp scatter<strong>in</strong>g. Both the scattered <strong>and</strong> the ejected proton suffer f<strong>in</strong>al-state <strong>in</strong>teractionswhile leav<strong>in</strong>g the nucleus. The scattered <strong>and</strong> the ejected proton are detected <strong>in</strong> co<strong>in</strong>cidence, whilethe residual nucleus rema<strong>in</strong>s unobserved.With regard to the treatment of the hard NN scatter<strong>in</strong>g part, essentially two methods exist.In the so-called cross-section factorized approximation [1, 4], the nucleon-nucleon scatter<strong>in</strong>g


Chapter 1. Introduction 3cross section enters as a multiplicative factor <strong>in</strong> the differential A(p, pN) cross section. Some resultsof exclusive A(p, 2p) measurements <strong>in</strong>terpreted with this cross-section factorized form canbe found <strong>in</strong> Refs. [5–10]. Inclusion of a sp<strong>in</strong> dependence <strong>in</strong> the description of the IFSI, however,breaks this factorization scheme. In that situation, an alternative technique can be used: the amplitudefactorized form of the cross section [11]. In this approach, the two-body NN <strong>in</strong>teractioncan be approximated by the <strong>in</strong>terpolation of phase shifts [12, 13] from free elastic NN scatter<strong>in</strong>g.Various phenomenological forms to fit the amplitudes have been proposed. Traditionally,the nucleon-nucleon scatter<strong>in</strong>g matrix has been parametrized <strong>in</strong> terms of five Lorentz <strong>in</strong>variants[14–23], a method usually dubbed as the IA1 model or the SPVAT (scalar, pseudoscalar,vector, axial vector, tensor) form of the NN scatter<strong>in</strong>g matrix. Cross-section calculations adopt<strong>in</strong>gthese five-term representations have been reported <strong>in</strong> Refs [24–32] <strong>and</strong> provide reasonablepredictions of A(p, pN) observables. However, such a five-term relativistic parametrization ofthe NN scatter<strong>in</strong>g matrix is <strong>in</strong>herently ambiguous [33]. Tjon <strong>and</strong> Wallace [34] have shownthat a complete expansion of the NN scatter<strong>in</strong>g matrix (commonly called the IA2 model) conta<strong>in</strong>s44 <strong>in</strong>dependent <strong>in</strong>variant amplitudes. To date, the sole calculations adopt<strong>in</strong>g this generalLorentz <strong>in</strong>variant representation have ignored all IFSI mechanisms [35].For the description of the target <strong>and</strong> residual nucleus, an <strong>in</strong>dependent-particle model (IPM)picture is customarily considered. Then, the <strong>in</strong>itial- <strong>and</strong> f<strong>in</strong>al-state (A + 1)-nucleon wave functionsare Slater determ<strong>in</strong>ants composed of s<strong>in</strong>gle-particle wave functions. These latter are solutionsto a one-body Schröd<strong>in</strong>ger or Dirac equation. In our approach, the relativistic boundstates<strong>in</strong>gle-particle wave functions are calculated <strong>in</strong> the Hartree approximation of the σ − ωmodel [36–38]. To take the IFSI <strong>in</strong>to account, the wave functions of the imp<strong>in</strong>g<strong>in</strong>g, scattered,<strong>and</strong> ejected nucleon are obta<strong>in</strong>ed <strong>in</strong> distorted-wave theory.The IFSI effects are typically computed <strong>in</strong> the distorted-wave impulse approximation(DWIA) framework [3, 4, 11, 39]. In a DWIA approach, the scatter<strong>in</strong>g wave functions of the<strong>in</strong>com<strong>in</strong>g <strong>and</strong> two outgo<strong>in</strong>g nucleons are generated by solv<strong>in</strong>g the Schröd<strong>in</strong>ger or Dirac equationwith complex optical potentials. Parametrizations for these optical potentials are usuallynot ga<strong>in</strong>ed from basic grounds, but are obta<strong>in</strong>ed by fitt<strong>in</strong>g elastic nucleon-nucleus scatter<strong>in</strong>gdata. Several sets of optical-potential parameters [9, 40–46] have been used <strong>in</strong> the descriptionof quasifree proton scatter<strong>in</strong>g off nuclei. In the past, both nonrelativistic <strong>and</strong> relativistic DWIAversions [5, 6, 8–10, 24–32, 47–51] have proven successful <strong>in</strong> predict<strong>in</strong>g A(p, pN) cross sectionsover a wide energy range (76–600 MeV) <strong>and</strong> for a whole scope of target nuclei. The comparisonof A(⃗p, 2p) analyz<strong>in</strong>g power calculations with data, on the other h<strong>and</strong>, clearly demonstratedthe superiority of the relativistic descriptions of the sp<strong>in</strong> observables [32, 52, 53]. Even withthe Dirac equation, however, predict<strong>in</strong>g the sp<strong>in</strong> observables for exclusive proton-knockoutreactions rema<strong>in</strong>s problematic [32, 52, 53].Most calculations for the exclusive A(p, pN) process addressed <strong>in</strong>cident proton k<strong>in</strong>etic energiesof a few hundred MeV. In this work, we aim at develop<strong>in</strong>g a formalism that extends toscatter<strong>in</strong>g <strong>in</strong> the GeV energy regime. As a matter of fact, the majority of DWIA frameworks


4rely on partial-wave expansions of the exact solution to the scatter<strong>in</strong>g problem, an approachwhich becomes <strong>in</strong>creas<strong>in</strong>gly cumbersome at higher energies. In this energy range, the eikonalapproximation (EA) [54, 55], which belongs to the class of high-energy semiclassical methods,offers a valid alternative for describ<strong>in</strong>g the IFSI. Nonrelativistic eikonal studies of the A(p, 2p)reaction used <strong>in</strong> comb<strong>in</strong>ation with optical potentials can be found <strong>in</strong> Refs. [1, 5, 7].In this work, we present a relativistic formulation of the eikonal approximation. Hereby,we extend a model developed by our research group, which was <strong>in</strong>itially designed for thedescription of exclusive electron-nucleus scatter<strong>in</strong>g processes [56–59], to deal with the IFSI <strong>in</strong>quasielastic proton-nucleus scatter<strong>in</strong>g. Our eikonal framework can accomodate both opticalpotentials <strong>and</strong> a multiple-scatter<strong>in</strong>g extension, Glauber theory. As such, the theoretical modelcan be formally applied <strong>in</strong> a wide energy range.Outl<strong>in</strong>eThe outl<strong>in</strong>e of this work is as follows.• In Chapter 2, a relativistic <strong>and</strong> cross-section factorized framework for comput<strong>in</strong>g quasielasticA(p, pN) observables at <strong>in</strong>termediate <strong>and</strong> high energies is presented. First, a factorizedexpression for the differential A(p, pN) cross section is derived. The rema<strong>in</strong>derof this chapter is devoted to the various eikonal methods to deal with the IFSI. Both anoptical-potential <strong>and</strong> a Glauber model are addressed. The optical-potential philosophy,which starts from a nucleon-nucleus picture, is particularly useful at lower nucleon energies.The Glauber method has nucleon-nucleon scatter<strong>in</strong>g as a start<strong>in</strong>g basis <strong>and</strong> is morenatural at higher energies. This versatility <strong>in</strong> describ<strong>in</strong>g the IFSI will allow us to describeA(p, pN) reactions <strong>in</strong> a wide energy range. The Glauber calculations are very dem<strong>and</strong><strong>in</strong>gas far as comput<strong>in</strong>g power is concerned s<strong>in</strong>ce they <strong>in</strong>volve multi-dimensional <strong>in</strong>tegralsover the positions of all spectator nucleons. In order to reduce computer time, an approximatedversion of our Glauber framework is <strong>in</strong>troduced. Its validity will be <strong>in</strong>vestigated<strong>in</strong> Chapter 4. F<strong>in</strong>ally, a second-order correction to the eikonal prescription is devised.This will enable us to probe the low-energy validity limit of the eikonal approach (seeChapter 5).• Chapter 3 starts with a detailed study of the properties of the IFSI factor <strong>in</strong> A(p, pN)reactions, a function where<strong>in</strong> the entire effect of the IFSI is conta<strong>in</strong>ed. Next, <strong>in</strong> order tovisualize which parts of the target nucleus are probed by the A(p, pN) reaction, the radial<strong>and</strong> polar-angle contributions to the cross sections are <strong>in</strong>vestigated. The results showthat A(p, pN) scatter<strong>in</strong>g events ma<strong>in</strong>ly take place at the nuclear surface. Further, ourcross-section calculations are compared to a wide variety of A(p, pN) experiments. Wepresent results for the target nuclei 4 He, 12 C, 16 O, <strong>and</strong> 40 Ca, <strong>and</strong> for different k<strong>in</strong>ematicalsett<strong>in</strong>gs.


Chapter 1. Introduction 5• The nuclear transparency extracted from A(p, 2p) scatter<strong>in</strong>g processes is the subject ofChapter 4. This observable gives a measure of the survival probability for the protons toenter <strong>and</strong> exit the nucleus without any further <strong>in</strong>teractions with the spectator nucleons.We start this chapter with a description of the features of high-momentum-transfer wideangleproton-proton scatter<strong>in</strong>g. There is a discussion of the different mechanisms play<strong>in</strong>ga role <strong>in</strong> this reaction: the quark-count<strong>in</strong>g <strong>and</strong> the L<strong>and</strong>shoff scatter<strong>in</strong>g. Next, a relativisticframework for comput<strong>in</strong>g the A(p, 2p) nuclear transparency is developed. Here, thenuclear filter<strong>in</strong>g mechanism is implemented as a possible explanation for the oscillatoryenergy dependence of the transparency. The phenomenon of color transparency (CT) istaken <strong>in</strong>to account as well. The special conditions necessary to observe CT are discussed<strong>and</strong> two exist<strong>in</strong>g models are used for its description. F<strong>in</strong>ally, results are shown for thetarget nuclei 7 Li, 12 C, 27 Al, <strong>and</strong> 63 Cu. The approximated, computationally less <strong>in</strong>tensiveversion of the Glauber framework is found to be sufficiently accurate for the calculationof the nuclear transparency. After <strong>in</strong>clud<strong>in</strong>g the nuclear filter<strong>in</strong>g <strong>and</strong> color transparencymechanisms, our calculations are <strong>in</strong> acceptable agreement with the data. As the outlookstates, however, there rema<strong>in</strong> a number of unanswered questions <strong>in</strong> connection with thenuclear transparency.• Due to its appeal<strong>in</strong>g potential, the eikonal approximation is widely used. However, it isnot clear down to which energies the eikonal approach is adequate. Therefore, <strong>in</strong> Chapter5, we turn our attention to lower energies <strong>and</strong> <strong>in</strong>vestigate the accuracy of the eikonalapproximation by tak<strong>in</strong>g second-order corrections <strong>in</strong>to account. We restrict ourselves toA(e, e ′ p) observables. First, the effect of the second-order corrections on the A(e, e ′ p) nucleartransparency is studied. Next, we focus on truly exclusive observables such as the<strong>in</strong>duced normal polarization, the left-right asymmetry, <strong>and</strong> the differential cross section.These observables are expected to be more sensitive to the details of the f<strong>in</strong>al-state <strong>in</strong>teractionsthan the nuclear transparency, which is an <strong>in</strong>tegrated quantity. As such, theeikonal approximation is put to an even more str<strong>in</strong>gent test. On the whole, the effect ofthe second-order corrections is found to be rather small down to four-momentum transfersQ 2 = 0.2 (GeV/c) 2 which corresponds with proton k<strong>in</strong>etic energies of slightly above100 MeV.• F<strong>in</strong>ally, Chapter 6 summarizes our f<strong>in</strong>d<strong>in</strong>gs <strong>and</strong> states our conclusions.


Chapter 2Relativistic Eikonal Description ofA(p, pN) ReactionsThe <strong>in</strong>terpretation of quasielastic nucleon-knockout reaction experiments heavily depends onthe availability of realistic models for the description of nucleon propagation through the nuclearmedium. With three nucleons subject to IFSI effects, this is certa<strong>in</strong>ly the case for A(p, pN)reactions. Unfortunately, there exists no uniform framework which can compute these attenuationeffects for both “low” <strong>and</strong> “high” energies. At lower energies, most theoretical workon IFSI is performed <strong>in</strong> the DWIA framework [3, 4, 11, 39]. The scatter<strong>in</strong>g state is obta<strong>in</strong>edas a partial-wave expansion <strong>and</strong> nucleon-nucleus optical potentials are used to estimate theeffect of the scatter<strong>in</strong>gs on the propagat<strong>in</strong>g particle. The validity of these DWIA models becomesquestionable for nucleon k<strong>in</strong>etic energies exceed<strong>in</strong>g 1 GeV. The first reason for this isthat the number of partial waves needed to reach convergence <strong>in</strong>creases rapidly with energy.Next, global parametrizations of optical potentials are usually restricted to nucleon k<strong>in</strong>etic energiesbelow 1 GeV. F<strong>in</strong>ally, the elementary nucleon-nucleon <strong>in</strong>teraction becomes highly <strong>in</strong>elasticabove 1 GeV, mak<strong>in</strong>g optical potentials unnatural <strong>in</strong> this high-energy regime. An alternative<strong>and</strong> economical description of IFSI is offered by the eikonal method [54, 55], which f<strong>in</strong>ds itsorig<strong>in</strong>s <strong>in</strong> optics [60, 61]. Be<strong>in</strong>g a semiclassical method, the EA becomes useful when the deBroglie wavelength λ dB = h/p of the particle with momentum p is small compared to the typicalrange a of the potential. This short-wavelength condition is equivalent to the requirementthat ka ≫ 1, with k = p/ the particle’s wavenumber. Fig. 2.1 depicts the dependence of thede Broglie wavelength of the knocked-out nucleon on the energy transfer ω. The de Brogliehcwavelength is approximately given by λ dB = √ , a relation obta<strong>in</strong>ed by neglect<strong>in</strong>g theω 2 +2M N ωmomentum of the residual nucleus relative to its rest mass. S<strong>in</strong>ce the <strong>in</strong>teraction range a istypically of the order of a few fm, the condition to be fulfilled can be written as ω ≫ 500 MeV.An additional assumption is that |V 0 |/E ≪ 1, with V 0 the typical strength of the potential <strong>and</strong>E the energy of the particle.In this chapter, we propose a relativistic <strong>and</strong> cross-section factorized formalism for comput<strong>in</strong>gexclusive A(p, pN) cross sections at <strong>in</strong>cident proton energies <strong>in</strong> the few hundred MeV to7


843λ dB(fm)2100 1 2 3 4ω (GeV)Figure 2.1 The de Broglie wavelength λ dB of the knocked-out nucleon as a function of the energy ωtransferred <strong>in</strong> an A(p, pN) reaction.GeV range. Our approach is founded on the impulse approximation <strong>and</strong> we adopt the eikonalapproximation to <strong>in</strong>corporate the IFSI effects. The eikonal formalism is implemented relativistically<strong>in</strong> comb<strong>in</strong>ation with optical potentials [9, 46], as well as with Glauber theory [62–64],which is a multiple-scatter<strong>in</strong>g extension of the EA. The two frameworks only differ <strong>in</strong> the waythey treat the IFSI.This chapter is organized as follows. In Sections 2.1.1 <strong>and</strong> 2.1.2, the factorized cross sectionis derived <strong>in</strong> the relativistic plane-wave impulse approximation (RPWIA) formalism. Anexpression for the differential A(p, pN) cross section when implement<strong>in</strong>g IFSI effects is formulated<strong>in</strong> Section 2.1.3. The different eikonal methods to deal with the IFSI are developed <strong>in</strong>the follow<strong>in</strong>g sections. In Section 2.2, the relativistic eikonal approximation is <strong>in</strong>troduced <strong>and</strong>complex optical potentials are implemented to describe the IFSI distortions. Next, the Glaubermultiple-scatter<strong>in</strong>g extension of the eikonal method is developed <strong>in</strong> Section 2.3. A numericallyless <strong>in</strong>tensive version of the Glauber framework is derived <strong>in</strong> Section 2.4, while Section 2.5 isdevoted to the derivation of a second-order correction to the relativistic optical model eikonalapproximation of Section 2.2.


2.1. Observables <strong>and</strong> K<strong>in</strong>ematics 10whereÔ (2) =A∑Ô (⃗r i , ⃗r j ) (2.5)i


Chapter 2. Relativistic Eikonal A(p, pN) Formalism 11The (A + 1)-body wave function <strong>in</strong> the f<strong>in</strong>al state readsΨ ⃗ k 1 ,m s1f , ⃗ k 2 ,m s2fA+1(⃗r 0 , ⃗r 1 , ⃗r 2 , ⃗r 3 )∣ ∣∣∣∣∣∣∣∣∣ φ ⃗k1 (⃗rm s1f 0 ) φ ⃗k2 (⃗rm s2f 0 ) φ α2 (⃗r 0 ) φ α3 (⃗r 0 )1 φ ⃗k1 (⃗rm= √ s1f 1 ) φ ⃗k2 (⃗rm s2f 1 ) φ α2 (⃗r 1 ) φ α3 (⃗r 1 )(A + 1)! φ ⃗k1 (⃗rm s1f 2 ) φ ⃗k2 (⃗rm s2f 2 ) φ α2 (⃗r 2 ) φ α3 (⃗r 2 ). (2.9)φ ⃗k1 (⃗rm s1f 3 ) φ ⃗k2 (⃗rm s2f 3 ) φ α2 (⃗r 3 ) φ α3 (⃗r 3 ) ∣Relative to the target nucleus ground state written <strong>in</strong> Eq. (2.7), the wave function of Eq. (2.9)refers to the situation whereby the struck proton resides <strong>in</strong> a state “α 1 ”, leav<strong>in</strong>g the residualA − 1 nucleus as a hole state <strong>in</strong> that particular s<strong>in</strong>gle-particle level. The outgo<strong>in</strong>g protons arerepresented by relativistic plane waves.S<strong>in</strong>ce both the <strong>in</strong>itial <strong>and</strong> the f<strong>in</strong>al wave functions are fully antisymmetrized, one can choosethe operator Ô(2) to act on two particular coord<strong>in</strong>ates (⃗r 0 <strong>and</strong> ⃗r 1 ). Without any loss of generality,the A(p, 2p) matrix element of Eq. (2.6) can be written as∫ ∫ ∫ ∫M (p,2p) A(A + 1) 1fi= d⃗r 0 d⃗r 1 d⃗r 2 d⃗r 32 (A + 1)!∑∑ ∑×k,l∈{ ⃗ k 1 m s1f , ⃗ k 2 m s2f}m,n∈{α 2 ,α 3 }× ɛ klmn ɛ opqr φ † k (⃗r 0) φ † l (⃗r 1) φ † m (⃗r 2 ) φ † n (⃗r 3 )o,p∈{⃗p 1 m s1i ,α 1 }∑q,r∈{α 2 ,α 3 }× Ô (⃗r 0, ⃗r 1 ) φ o (⃗r 0 ) φ p (⃗r 1 ) φ q (⃗r 2 ) φ r (⃗r 3 ) , (2.10)with ɛ ijkl the Levi-Civita tensor. In the RPWIA,∫ ∫ ∫d⃗r 0 d⃗r 1 d⃗r 2 φ † k (⃗r 0) φ † l (⃗r 1) φ † m (⃗r 2 ) Ô (⃗r 0, ⃗r 1 ) φ o (⃗r 0 ) φ p (⃗r 1 ) φ q (⃗r 2 )∫ ∫ ∫= δ mq d⃗r 0 d⃗r 1 d⃗r 2 φ † k (⃗r 0) φ † l (⃗r 1)× Ô (⃗r 0, ⃗r 1 ) φ o (⃗r 0 ) φ p (⃗r 1 ) |φ q (⃗r 2 )| 2 . (2.11)Insert<strong>in</strong>g this expression <strong>in</strong> Eq. (2.10), one obta<strong>in</strong>s∫ ∫ ∫M (p,2p) A(A + 1) 1fi= d⃗r 0 d⃗r 12 (A + 1)!∑∑×k,l∈{ ⃗ k 1 m s1f , ⃗ k 2 m s2f}o,p∈{⃗p 1 m s1i ,α 1 }d⃗r 2∫d⃗r 3∑m,n∈{α 2 ,α 3 }× ɛ klmn ɛ opmn φ † k (⃗r 0) φ † l (⃗r 1) |φ m (⃗r 2 )| 2 |φ n (⃗r 3 )| 2× Ô (⃗r 0, ⃗r 1 ) φ o (⃗r 0 ) φ p (⃗r 1 ) . (2.12)There are (A − 1)! possible choices (permutations) for the <strong>in</strong>dices m, n, . . . , all giv<strong>in</strong>g the samecontribution to the matrix element. Accord<strong>in</strong>gly, the above expression can be rewritten asM (p,2p)fi= 1 ∫ ∫ ∫ ∫∑∑d⃗r 0 d⃗r 1 d⃗r 2 d⃗r 32k,l∈{ ⃗ k 1 m s1f , ⃗ k 2 m s2f}× ɛ klα2 α 3ɛ opα2 α 3φ † k (⃗r 0) φ † l (⃗r 1) |φ α2 (⃗r 2 )| 2 |φ α3 (⃗r 3 )| 2o,p∈{⃗p 1 m s1i ,α 1 }× Ô (⃗r 0, ⃗r 1 ) φ o (⃗r 0 ) φ p (⃗r 1 ) . (2.13)


2.1. Observables <strong>and</strong> K<strong>in</strong>ematics 12Because the bound-state wave functions are normalized to unity ( ∫ d⃗r |φ α (⃗r)| 2Ô (⃗r 0 , ⃗r 1 ) = Ô (⃗r 1, ⃗r 0 ), the matrix element can be further simplified to∫ ∫)M (p,2p)fi= d⃗r 0 d⃗r 1(φ † ⃗(⃗r 0 ) φ † k1 m s1f⃗(⃗r 1 ) − φ † k2 m s2f⃗(⃗r 0 ) φ † k2 m s2f⃗(⃗r 1 )k1 m s1f= 1) <strong>and</strong>× Ô (⃗r 0, ⃗r 1 ) φ ⃗p1ms1i (⃗r 0 ) φ α1 (⃗r 1 ) , (2.14)<strong>in</strong>clud<strong>in</strong>g a direct <strong>and</strong> an exchange term.Substitution of the general form of the scatter<strong>in</strong>g operator∫d⃗pÔ (⃗r 0 , ⃗r 1 ) = ei⃗p·(⃗r 1−⃗r 0 ) ̂F (⃗p) , (2.15)(2π) 3where ̂F (⃗p) is the NN scatter<strong>in</strong>g amplitude <strong>in</strong> momentum space, <strong>in</strong> the above expression (2.14)leads to∫∫d⃗p(2π) 3d⃗r 0∫d⃗r 1 e −i⃗ k 1·⃗r 0u † ( ⃗ k 1 , m s1f ) e −i⃗ k 2·⃗r 1u † ( ⃗ k 2 , m s2f ) e i⃗p·(⃗r 1−⃗r 0 )× ̂F (⃗p) e i⃗p 1·⃗r 0u(⃗p 1 , m s1i ) φ α1 (⃗r 1 )= u † ( ⃗ k 1 , m s1f ) u † ( ⃗ k 2 , m s2f ) ̂F(⃗p 1 − ⃗ )k 1 u(⃗p 1 , m s1i ) φ α1 (⃗p m ) (2.16)for the direct term <strong>and</strong> a similar expression for the exchange term. Here, φ α (⃗p) is the relativisticwave function for the bound nucleon <strong>in</strong> momentum space (for details see Appendix B) <strong>and</strong>⃗p m = ⃗ k 1 + ⃗ k 2 − ⃗p 1 is the miss<strong>in</strong>g momentum. In order to arrive at a cross-section factorizedexpression for Eq. (2.2), the quasielastic off-shell proton-proton scatter<strong>in</strong>g matrix element willbe related to the free on-shell proton-proton cross section.completeness relation∑For this purpose, we <strong>in</strong>sert them s[u(⃗p m , m s ) ū(⃗p m , m s ) − v(⃗p m , m s ) ¯v(⃗p m , m s )] = 1 (2.17)<strong>in</strong>fi= u † ( ⃗ k 1 , m s1f ) u † ( ⃗ k 2 , m s2f ) ̂F(⃗p 1 − ⃗ )k 1 u(⃗p 1 , m s1i ) φ α1 (⃗p m )(− ⃗k1 m s1f ↔ ⃗ )k 2 m s2f , (2.18)M (p,2p)<strong>and</strong> obta<strong>in</strong> the follow<strong>in</strong>g expression for the matrix element:M (p,2p)fi= ∑ ( )M ppfiū(⃗p m , m s ) φ α1 (⃗p m )mm s1i ,m s,m s1f ,m s2f sHere, M ppfi(M ppfi− negative-energy projection term . (2.19)is the matrix element for free pp scatter<strong>in</strong>g)m s1i ,m s2i ,m s1f ,m s2f= u † ( ⃗ k 1 , m s1f ) u † ( ⃗ k 2 , m s2f ) ̂F(⃗p 1 − ⃗ )k 1 u(⃗p 1 , m s1i ) u(⃗p 2 , m s2i )(− ⃗k1 m s1f ↔ ⃗ )k 2 m s2f . (2.20)


Chapter 2. Relativistic Eikonal A(p, pN) Formalism 13At the level of the differential cross section, the decomposition of Eq. (2.19) results <strong>in</strong> threeterms: two terms proportional to the square of the positive-energy ū φ α <strong>and</strong> negative-energy¯v φ α projection, respectively, <strong>and</strong> a cross term conta<strong>in</strong><strong>in</strong>g the product of the positive- <strong>and</strong>negative-energy projections. Us<strong>in</strong>g expression (B.7) for the relativistic bound-nucleon wavefunction <strong>in</strong> momentum space, the ū φ α <strong>and</strong> ¯v φ α contractions reduce to [66]ū(⃗p, m s ) φ α (⃗p) = (−i) l (2π) 3/2 ˜α nκ (p) χ † 12 ms Y κm (Ω p ) , (2.21)¯v(⃗p, m s ) φ α (⃗p) = (−i) l (2π) 3/2 ˜βnκ (p) χ † 12 ms Y −κm (Ω p ) , (2.22)where χ † 1 Y ±κm <strong>in</strong>dicates the sp<strong>in</strong> projection of the sp<strong>in</strong> spherical harmonic Y ±κm (Ω p ) on a2 mssp<strong>in</strong> state χ 1 <strong>and</strong> the radial functions <strong>in</strong> momentum space ˜α 2 ms nκ <strong>and</strong> ˜β nκ are given by˜α nκ (p) =(√Ē + Mpg nκ (p) −2M ppĒ + M pS κ f nκ (p)), (2.23)()√Ē + Mp p˜β nκ (p) = g nκ (p) − S κ f nκ (p) , (2.24)2M p Ē + M pwith g nκ <strong>and</strong> f nκ the Bessel transforms of the upper <strong>and</strong> lower√radial functions of the boundnucleonwave function <strong>in</strong> coord<strong>in</strong>ate space (see Eq. (B.8)), Ē = p 2 + Mp 2 <strong>and</strong> S κ = κ/ |κ|. Assuch, the contribution to the differential cross section from positive- (negative-)energy projectionsonly is proportional to the momentum distribution component |˜α nκ (p m )| 2 (| ˜β nκ (p m )| 2 ),while the cross term is proportional to 2|˜α nκ (p m ) ˜β nκ (p m )|.In Fig. 2.2 the projection componentsof the momentum distribution are shown for the 1p 1/2 shell of 16 O. The component|˜α nκ (p m )| 2 clearly dom<strong>in</strong>ates <strong>in</strong> the region p m ≤ 300 MeV/c. In fact, <strong>in</strong> the maximum atp m ≈ 100 MeV/c, it is one order of magnitude larger than the 2|˜α nκ (p m ) ˜β nκ (p m )| component<strong>and</strong> more than two orders of magnitude larger than the | ˜β nκ (p m )| 2 component. At high miss<strong>in</strong>gmomenta p m > 300 MeV/c, the situation is clearly different. In this p m region, the valuesof the components 2|˜α nκ (p m ) ˜β nκ (p m )| <strong>and</strong> | ˜β nκ (p m )| 2 are similar to or even larger than that of|˜α nκ (p m )| 2 .Ow<strong>in</strong>g to the presence of the negative-energy projection term, factorization breaks down,even when IFSI are disregarded. To recover factorization, the negative-energy projection termis neglected <strong>in</strong> the rema<strong>in</strong>der of this workM (p,2p)fi≈ ∑ ( )M ppfimm s1i ,m s,m s1f ,m s2f sū(⃗p m , m s ) φ α1 (⃗p m ) . (2.25)In view of the relative importance of the |˜α nκ (p m )| 2 term, this approximation may be expectedto be adequate up to miss<strong>in</strong>g momenta p m ≈ 300 MeV/c. Upon squar<strong>in</strong>g Eq. (2.25), the pp <strong>and</strong>nuclear bound-state parts get coupled by the summation over the <strong>in</strong>termediate sp<strong>in</strong>s m s <strong>and</strong>m ′ s:∣∣M (p,2p)fi∣ 2 ≈ ∑m s,m ′ s(M ppfi) ∗m s1i ,m s,m s1f ,m s2f(M ppfi)m s1i ,m ′ s,m s1f ,m s2f× (ū(⃗p m , m s ) φ α1 (⃗p m )) ∗ ū(⃗p m , m ′ s) φ α1 (⃗p m ) . (2.26)


2.1. Observables <strong>and</strong> K<strong>in</strong>ematics 14Momentum distribution components (fm 3 )10110 -110 -210 -310 -410 -510 -610 -7|α ~ nκ (p)|22|α ~ nκ (p) β~ nκ (p)||β ~ nκ (p)|210 -80 50 100 150 200 250 300 350 400 450 500p (MeV/c)Figure 2.2 The projection components of the momentum distribution for the 1p 1/2 shell of 16 O. The solid(dashed) l<strong>in</strong>e shows the contribution from positive- (negative-)energy projections only. The cross term,which <strong>in</strong>volves both positive- <strong>and</strong> negative-energy projections, is represented by the dot-dashed l<strong>in</strong>e.After summation over m, the struck nucleon’s generalized angular momentum quantum number,the square of ū(⃗p m , m s ) φ α1 (⃗p m ) yields a δ msm ′ s , i.e., becomes diagonal <strong>in</strong> m s. Thereby, useis made of the identity∑( ) ∗χ † 1 Y κm χ †2 ms 1 Y κm = 2j + 12 m′ s 8π δ m sm ′ . (2.27)smThis leads to the decoupl<strong>in</strong>g between the pp scatter<strong>in</strong>g <strong>and</strong> the bound-state part <strong>in</strong> the matrixelement:∑∣if∣M (p,2p)fi∣ 2 ≈ (2π) 3 2j + 1 |˜α nκ (p m )| 28π× ∑ m s1i ,m s1f ,m s2f∑m s∣ ∣∣∣ (M ppfi)m s1i ,m s,m s1f ,m s2f∣ ∣∣∣2. (2.28)The last factor <strong>in</strong> Eq. (2.28) can be related to the free pp scatter<strong>in</strong>g center-of-mass (c.m.) crosssection( ) dσppdΩc.m.= M p4 1∑∣ ∣∣∣ ((2π) 2 s 2∑m s1i ,m s1f ,m s2fm sM ppfi∣ ∣∣∣2, (2.29))m s1i ,m s,m s1f ,m s2f


Chapter 2. Relativistic Eikonal A(p, pN) Formalism 15so that the RPWIA differential A(p, 2p) cross section of Eq. (2.2) can be written <strong>in</strong> the crosssectionfactorized form(d 5 ) RPWIAσ≈ sM A−1 k 1 k 2f −1 2j + 1recdE k1 dΩ 1 dΩ 2 M p M A p 1 4πHere, s is the M<strong>and</strong>elstam variable for the pp scatter<strong>in</strong>g.(dσ pNdΩ( )|˜α nκ (p m )| 2 dσppdΩc.m.. (2.30)In)the numerical calculations presented <strong>in</strong> Chapter 3, the free proton-nucleon cross sectionis obta<strong>in</strong>ed from the SAID code [12, 13] at an effective laboratory k<strong>in</strong>etic energy ofc.m.T efflab = s − 4M 2 p2M p(2.31)<strong>and</strong> a c.m. scatter<strong>in</strong>g angle θ c.m. given bycos θ c.m. =t − u√ (s− 4M 2 p) [ (4M 2 p −t−u) 2s− 4M 2 p] , (2.32)with s, t, <strong>and</strong> u the M<strong>and</strong>elstam variables for the pp scatter<strong>in</strong>g. For the nuclear transparencycalculations of Chapter 4, the phenomenological parametrization of Ref. [67], which is moresuitable at large s (s ≥ 7 GeV 2 ) <strong>and</strong> c.m. scatter<strong>in</strong>g angle θ c.m. ≈ 90 ◦ , is used (see Eq. (4.8)).2.1.3 Treatment of the IFSI: Factorization Assumption <strong>and</strong> the Distorted MomentumDistributionThe factorized RPWIA result of Eq. (2.30) adopts an oversimplified description of the reactionmechanism. The momentum distribution 2j+14π|˜α nκ (p m )| 2 , which represents the probability off<strong>in</strong>d<strong>in</strong>g a proton <strong>in</strong> the target nucleus with miss<strong>in</strong>g momentum ⃗p m , is modified by the scatter<strong>in</strong>gsof the <strong>in</strong>com<strong>in</strong>g <strong>and</strong> outgo<strong>in</strong>g protons <strong>in</strong> the nucleus. Therefore, it is necessary to <strong>in</strong>corporatethe effects of these IFSI <strong>in</strong> the model.In this section, the differential A(p, 2p) cross section is written <strong>in</strong> a factorized form tak<strong>in</strong>gIFSI effects <strong>in</strong>to account. The relativistic eikonal methods used for deal<strong>in</strong>g with the IFSI effectswill be discussed <strong>in</strong> depth <strong>in</strong> the forthcom<strong>in</strong>g sections. Two methods will be used. The relativisticoptical model eikonal approximation (ROMEA) is the subject of Section 2.2, whereas therelativistic multiple-scatter<strong>in</strong>g Glauber approximation (RMSGA) is discussed <strong>in</strong> Section 2.3.In both versions of the relativistic eikonal framework for A(p, pN) reactions (ROMEA <strong>and</strong>RMSGA), the antisymmetrized <strong>in</strong>itial- <strong>and</strong> f<strong>in</strong>al-state (A + 1)-body wave functions,Ψ ⃗p 1,m s1i ,gsA+1(⃗r 0 , ⃗r 1 , . . . , ⃗r A )[= ÂŜ p1 (⃗r 0 , ⃗r 2 , . . . , ⃗r A ) e i⃗p 1·⃗r 0u(⃗p 1 , m s1i ) Ψ gsA (⃗r 1, ⃗r 2 , . . . , ⃗r A )](2.33)<strong>and</strong>Ψ ⃗ k 1 ,m s1f , ⃗ k 2 ,m s2fA+1(⃗r 0 , ⃗r 1 , . . . , ⃗r A ) = Â[Ŝ † k1 (⃗r 0, ⃗r 2 , . . . , ⃗r A ) e i⃗ k 1·⃗r 0u( ⃗ k 1 , m s1f )× Ŝ† k2 (⃗r 1, ⃗r 2 , . . . , ⃗r A ) e i⃗ k 2·⃗r 1u( ⃗ k 2 , m s2f ) Ψ J R M RA−1(⃗r 2 , . . . , ⃗r A )], (2.34)


2.1. Observables <strong>and</strong> K<strong>in</strong>ematics 16differ from their RPWIA counterparts of Eqs. (2.7) <strong>and</strong> (2.9) through the presence of the operatorsŜp1, Ŝk1, <strong>and</strong> Ŝk2. These def<strong>in</strong>e the accumulated effect of all <strong>in</strong>teractions that the <strong>in</strong>com<strong>in</strong>g<strong>and</strong> emerg<strong>in</strong>g protons undergo <strong>in</strong> their way <strong>in</strong>to <strong>and</strong> out of the target nucleus.S<strong>in</strong>ce the IFSI violate factorization, some additional approximations are <strong>in</strong> order. First, onlycentral IFSI are considered. In both proton-nucleus <strong>and</strong> A(e, e ′ p) scatter<strong>in</strong>g calculations (see,for example, Refs. [59,68]), it has been a common <strong>and</strong> successful practice to neglect sp<strong>in</strong> effects.Hence, we expect a sp<strong>in</strong>less treatment of the IFSI to be quite reasonable for the calculationof A(p, pN) cross sections as well, especially at higher energies, s<strong>in</strong>ce the contribution fromsp<strong>in</strong>-dependent terms decreases rapidly with energy. Of course, sp<strong>in</strong> effects might play a moreprom<strong>in</strong>ent role <strong>in</strong> the description of A(p, 2p) sp<strong>in</strong> observables, but these observables will notbe discussed <strong>in</strong> this work. Further, the zero-range approximation is adopted for the hard NN<strong>in</strong>teraction, allow<strong>in</strong>g one to replace the coord<strong>in</strong>ates of the two <strong>in</strong>teract<strong>in</strong>g protons (⃗r 0 <strong>and</strong> ⃗r 1 )by one s<strong>in</strong>gle collision po<strong>in</strong>t <strong>in</strong> the distort<strong>in</strong>g functions Ŝp1, Ŝ k1 , <strong>and</strong> Ŝk2. This leads to thedistorted momentum-space wave function∫φ D α 1(⃗p m ) = d⃗r e −i⃗pm·⃗r φ α1 (⃗r) ŜIFSI (⃗r) , (2.35)similar to Eq. (B.7), but with the additional IFSI factor∫ ∫Ŝ IFSI (⃗r) = d⃗r 2 · · · d⃗r A |φ α2 (⃗r 2 )| 2 · · · |φ αA (⃗r A )| 2 Ŝ k1 (⃗r, ⃗r 2 , . . . , ⃗r A )account<strong>in</strong>g for the soft IFSI effects.× Ŝk2 (⃗r, ⃗r 2 , . . . , ⃗r A ) Ŝp1 (⃗r, ⃗r 2 , . . . , ⃗r A ) (2.36)Now, along the l<strong>in</strong>es of [69], it is natural to def<strong>in</strong>e a distorted wave amplitudeψ D (⃗p m ) = ū(⃗p m , m s ) φ D α 1(⃗p m ) , (2.37)so that the distorted momentum distribution is given by the square of this amplitude,ρ D (⃗p m ) = 1 ∑ ∑ ∣ (2π) 3 ψ D (⃗p m ) ∣ 2 . (2.38)m m sThis distorted momentum distribution has the follow<strong>in</strong>g properties. First, it takes <strong>in</strong>to accountthe distortions for the <strong>in</strong>com<strong>in</strong>g <strong>and</strong> outgo<strong>in</strong>g protons. Second, it reduces to the plane-wavemomentum distribution 2j+14π |˜α nκ (p m )| 2 <strong>in</strong> the plane-wave limit when assum<strong>in</strong>g that φ α1 (⃗p m )satisfies the relation⃗σ · ⃗pĒ + M pφ u = φ d (2.39)between the upper <strong>and</strong> lower components.Us<strong>in</strong>g the ansatz (2.38) for the distorted momentum distribution, the differential A(p, 2p)cross section can be cast <strong>in</strong> the form(d 5 ) Dσ≈ sM ( )A−1 k 1 k 2dσfrec −1 ρ D pp(⃗p m )dE k1 dΩ 1 dΩ 2 M p M A p 1dΩc.m.. (2.40)It differs from the RPWIA expression (2.30) through the <strong>in</strong>troduction of a “distorted” momentumdistribution ρ D .


Chapter 2. Relativistic Eikonal A(p, pN) Formalism 172.2 Relativistic Optical Model Eikonal Approximation2.2.1 Nucleon-Nucleus Scatter<strong>in</strong>gFollow<strong>in</strong>g the discussion of Refs. [56,70], √we consider the time-<strong>in</strong>dependent Dirac equation fora particle with relativistic energy E = k 2 + MN 2 <strong>and</strong> sp<strong>in</strong> state ∣ 12 m s〉subject to a sphericalLorentz scalar V s (r) <strong>and</strong> vector potential V v (r)Ĥ Ψ (+)⃗ k,ms(⃗r) = [⃗α · ˆ⃗p + βM N + βV s (r) + V v (r)] Ψ (+)⃗ k,ms(⃗r) = E Ψ (+)⃗ k,ms(⃗r) , (2.41)where Ψ (+)⃗ k,ms(⃗r) is the unbound (scattered) Dirac state <strong>and</strong> ˆ⃗p represents the impulse operator.The <strong>in</strong>fluence of the nuclear medium on the particle is twofold: the scalar potential V s (r) shiftsthe particle mass to an effective value, while the vector potential V v (r) affects the energy term.The scatter<strong>in</strong>g wave function Ψ (+)⃗ k,ms(⃗r) is decomposed <strong>in</strong> an upper <strong>and</strong> a lower component,u (+)⃗ k,ms(⃗r) <strong>and</strong> w (+)⃗ k,ms(⃗r). Some straightforward manipulations lead to a Schröd<strong>in</strong>ger-like equationfor the upper component][− ∇2+ V c (r) + V so (r) (⃗σ ·2M ⃗L − i⃗r · ˆ⃗p) u (+)⃗(⃗r) =Nk,mswhile the lower component is related to the upper one throughw (+)⃗ k,ms(⃗r) =k22M Nu (+)⃗ k,ms(⃗r) , (2.42)1E + M N + V s (r) − V v (r) ⃗σ · ˆ⃗p u (+)⃗ k,ms(⃗r) . (2.43)Here, the central <strong>and</strong> sp<strong>in</strong>-orbit potentials V c (r) <strong>and</strong> V so (r) are def<strong>in</strong>ed asV c (r) = V s (r) + E V v (r) + V s 2 (r) − Vv 2 (r),M N 2M N11 dV so (r) =2M N [E + M N + V s (r) − V v (r)] r dr [V v(r) − V s (r)] . (2.44)So far, no approximations have been made. In the relativistic DWIA frameworks [24,25,29,30], the scatter<strong>in</strong>g wave function is exp<strong>and</strong>ed <strong>in</strong> partial waves√Ψ (+)E + MN ∑⃗(⃗r) = 4πi l 1〈lm l k,ms 2M N2 m s|jm〉 Ylm ∗l(Ω k ) Ψ m κ (⃗r) , (2.45)κmm lwhere Ψ m κ (⃗r) are four-sp<strong>in</strong>ors of the same form as the bound-state wave functions of Eq. (B.4),<strong>and</strong> Eq. (2.41) is solved numerically us<strong>in</strong>g optical potentials. This partial-wave procedure becomesimpractical as the energy <strong>in</strong>creases. Therefore, at higher energies, the Schröd<strong>in</strong>ger-typeequation (2.42) is solved <strong>in</strong> the eikonal approximation [56, 70].Follow<strong>in</strong>g the method outl<strong>in</strong>ed <strong>in</strong> Ref. [70], the average momentum ⃗ K <strong>and</strong> the momentumtransfer ⃗ ∆ which occur dur<strong>in</strong>g the nucleon-nucleus collision, are def<strong>in</strong>ed <strong>in</strong> terms of thenucleon’s <strong>in</strong>itial ( ⃗ k i ) <strong>and</strong> f<strong>in</strong>al momentum ( ⃗ k f )⃗K = ⃗ k i + ⃗ k f,2⃗∆ = ⃗ k i − ⃗ k f . (2.46)


2.2. Relativistic Optical Model Eikonal Approximation 18The EA is essentially a small-angle approximation (∆/k i ≪ 1) <strong>and</strong> hence, the follow<strong>in</strong>g operatorialapproximation is made [71]ˆp 2 = [(ˆ⃗p − ⃗ K) + ⃗ K] 2 ≃ 2 ⃗ K · ˆ⃗p − K 2 . (2.47)Through this, the equation for the upper component becomes l<strong>in</strong>ear <strong>in</strong> the momentum operator:[ {⃗K · ˆ⃗p − K 2 + M N V c (r) + V so (r) (⃗σ · ⃗r × K ⃗ − i⃗r · ⃗K)}]u (+)⃗(⃗r) = 0 , (2.48)k,mswhere the momentum operators <strong>in</strong> the sp<strong>in</strong>-orbit (V so (r) (⃗σ · ⃗r × ˆ⃗p)) <strong>and</strong> Darw<strong>in</strong> terms(V so (r) (−i⃗r · ˆ⃗p)) have been replaced by the average momentum ⃗ K. In the EA, the follow<strong>in</strong>gansatz is postulated for the upper component of the scatter<strong>in</strong>g wave function:u (+)⃗(⃗r) ≡ N e i S b eik (⃗r) e i⃗k·⃗r χ 1 , (2.49)k,ms ms2with N a normalization factor. Insert<strong>in</strong>g this expression <strong>in</strong>to Eq. (2.48) yields a differentialequation for the eikonal phase Ŝeik(⃗r), which is an operator <strong>in</strong> sp<strong>in</strong> space. Def<strong>in</strong><strong>in</strong>g the z axisalong the direction of the average momentum ⃗ K, this eikonal phase can be written <strong>in</strong> the <strong>in</strong>tegralform (⃗r ≡ ( ⃗ b, z))i Ŝeik( ⃗ b, z) = −i M NK∫ z−∞{ dz ′ V c ( ⃗ b, z ′ ) + V so ( ⃗ b, z ′ ) (⃗σ ·⃗b × K ⃗ }− iKz ′ )In the relativistic eikonal limit, the scatter<strong>in</strong>g wave function takes on the form√ []Ψ (+) E + MN1⃗(⃗r) =k,ms 2M N ⃗σ · ˆ⃗p1E+M N +V s(r)−V v(r). (2.50)e i S b eik (⃗r) e i⃗k·⃗r χ 1 . (2.51)ms2It is normalized such that it co<strong>in</strong>cides with the relativistic plane wave (2.8) when ⃗r → −∞.Eq. (2.51) differs from the plane-wave solutions <strong>in</strong> two respects. First, the lower component isdynamically enhanced due to the comb<strong>in</strong>ation of the scalar <strong>and</strong> vector potentials V s − V v < 0.Second, the eikonal phase e i b S eik (⃗r) describes the <strong>in</strong>teractions of the nucleon with the nucleus viapotential scatter<strong>in</strong>g. Furthermore, one can identify two primary relativistic effects: the Darw<strong>in</strong>term V so ( ⃗ b, z ′ ) (−iKz ′ ) <strong>in</strong> Eq. (2.50) <strong>and</strong> the previously mentioned dynamical enhancement ofthe lower component. The EA reproduces the exact partial-wave results well <strong>in</strong> <strong>in</strong>termediateenergyproton-nucleus scatter<strong>in</strong>g (T p ≈ 500 MeV) [70] <strong>and</strong> has also been successfully applied<strong>in</strong> A(e, e ′ p) scatter<strong>in</strong>g [56, 57, 72–76].The eikonal phase (2.50) is determ<strong>in</strong>ed by perform<strong>in</strong>g a straight l<strong>in</strong>e <strong>in</strong>tegration along thedirection of ⃗ K. A more accurate computation of the scatter<strong>in</strong>g wave function would be obta<strong>in</strong>edby calculat<strong>in</strong>g its phase along the actual curved classical trajectory. Therefore, the useof the EA is only justified for small-angle scatter<strong>in</strong>g. Fortunately, high-energy proton-nucleuscollisions like the soft IFSI <strong>in</strong> A(p, pN) reactions are diffractive <strong>and</strong> extremely forward peaked.The scatter<strong>in</strong>g wave function Ψ (+)⃗ k,ms(⃗r) of Eq. (2.51) satisfies outgo<strong>in</strong>g boundary conditions<strong>and</strong> can be used to describe the <strong>in</strong>itial-state <strong>in</strong>teractions (ISI) of the imp<strong>in</strong>g<strong>in</strong>g proton. For


Chapter 2. Relativistic Eikonal A(p, pN) Formalism 19the description of the f<strong>in</strong>al-state <strong>in</strong>teractions (FSI), however, <strong>in</strong>com<strong>in</strong>g boundary conditionsare appropriate. Accord<strong>in</strong>g to st<strong>and</strong>ard distorted-wave theory [77], the correspond<strong>in</strong>g wavefunction Ψ (−)⃗(⃗r) is related to Ψ (+)k,ms ⃗(⃗r) by time reversal. Under time reversal, the follow<strong>in</strong>gk,mstransformations occur:t → −t ,⃗r → ⃗r ,ˆ⃗p → −ˆ⃗p ,⃗σ → −⃗σ ,⃗L → −L ⃗ ,c → c ∗ ,(2.52)where the last l<strong>in</strong>e <strong>in</strong>dicates that all complex numbers are transformed <strong>in</strong>to their complex conjugates.Thus, Ψ (−)⃗ k,ms(⃗r) satisfies Eqs. (2.41)–(2.43) if the potentials V s (r), V v (r), V c (r), <strong>and</strong> V so (r)are replaced by their complex conjugates. The eikonal solution satisfy<strong>in</strong>g <strong>in</strong>com<strong>in</strong>g boundaryconditions then takes the formΨ (−)⃗ k,ms(⃗r) =√ [E + MN2M N(× exp i M NK(× exp i ⃗ )k · ⃗r11E+M N +Vs ∗ (r)−Vv ∗ (r)∫ +∞z⃗σ · ˆ⃗p]dz ′ { V ∗c ( ⃗ b, z ′ ) + V ∗so( ⃗ b, z ′ ) (⃗σ ·⃗b × ⃗ K − iKz ′ )} )χ 1 , (2.53)ms2or, <strong>in</strong> the conjugate “bra” form <strong>in</strong> which the outgo<strong>in</strong>g wave functions appear <strong>in</strong> the A(p, 2p)matrix element(†√Ψ (−)E + MN(⃗(⃗r))=χ † 1 exp −i k,ms 2M ⃗ )k · ⃗rN 2(ms× exp −i M ∫ +∞ { Ndz ′ V c (K⃗ b, z ′ ) + V so ( ⃗ b, z ′ ) (⃗σ ·⃗b × ⃗ } )K + iKz ′ )z[× 1 −⃗σ · ˆ⃗p]1E+M N +V s(r)−V v(r). (2.54)2.2.2 ROMEA for A(p, pN) ReactionsIn evaluat<strong>in</strong>g the IFSI effects <strong>in</strong> our A(p, pN) calculations, some approximations are <strong>in</strong>troduced.First, the dynamical enhancement of the lower components of the scatter<strong>in</strong>g wave functions(2.51) is neglected <strong>in</strong> our factorized approach, the so-called noSV approximation. For smallmomenta, the lower components play a m<strong>in</strong>or role with respect to the upper ones, due to thefactor ˆ⃗p/(E + M N + V s (r) − V v (r)); while at higher momenta, V s (r) − V v (r) can be disregarded<strong>in</strong> comparison with E + M N . As such, the effect of the dynamical enhancement is not expectedto be important for the A(p, pN) cross sections. Next, the average momentum ⃗ K is approximatedby the asymptotic momenta of the imp<strong>in</strong>g<strong>in</strong>g <strong>and</strong> outgo<strong>in</strong>g nucleons (⃗p 1 , ⃗ k 1 , <strong>and</strong> ⃗ k 2 ).This is allowed with<strong>in</strong> the small-angle restriction of the EA. Further, <strong>in</strong> the calculation of thescatter<strong>in</strong>g states, the impulse operator ˆ⃗p is replaced by the asymptotic momenta of the nucleons.In literature, this is usually referred to as the effective momentum approximation (EMA),


2.2. Relativistic Optical Model Eikonal Approximation 20<strong>in</strong> which the momentum operators that appear <strong>in</strong> sp<strong>in</strong>or-distortion operators are replaced byasymptotic k<strong>in</strong>ematics [78]. As mentioned before, the sp<strong>in</strong>-orbit potential V so is also omitted.As a result, the effects of the <strong>in</strong>teractions of the <strong>in</strong>com<strong>in</strong>g <strong>and</strong> outgo<strong>in</strong>g nucleons withthe residual nucleus are implemented <strong>in</strong> the distorted momentum-space wave function ofEq. (2.35) through the follow<strong>in</strong>g phase factors:Mp R zp1−iŜ p1 (⃗r) = ep 1−∞ dz Vc(⃗ b p1 ,z) ,(2.55a)Mp R−i +∞kŜ k1 (⃗r) = e 1 z dz ′ V c( ⃗ b k1 ,z ′ ) k1 , (2.55b)Ŝ k2 (⃗r) = e −i M Nk2R +∞z k2dz ′′ V c( ⃗ b k2 ,z ′′ ), (2.55c)with the z-axes of the different coord<strong>in</strong>ate systems ly<strong>in</strong>g along the trajectories of the respectiveparticles (z along the direction of the <strong>in</strong>com<strong>in</strong>g proton ⃗p 1 , z ′ along the trajectory of the scatteredproton ⃗ k 1 , <strong>and</strong> z ′′ along the path of the ejected nucleon ⃗ k 2 ); <strong>and</strong> ( ⃗ b p1 , z p1 ), ( ⃗ b k1 , z k1 ), <strong>and</strong> ( ⃗ b k2 , z k2 )are the coord<strong>in</strong>ates of the collision po<strong>in</strong>t ⃗r <strong>in</strong> the respective coord<strong>in</strong>ate systems. The geometryof the A(p, pN) scatter<strong>in</strong>g process is illustrated <strong>in</strong> Fig. 2.3. The <strong>in</strong>tegration limits guarantee thatthe <strong>in</strong>com<strong>in</strong>g proton only undergoes ISI up to the po<strong>in</strong>t where the hard NN collision occurs,<strong>and</strong> the outgo<strong>in</strong>g nucleons are only subject to FSI after this hard collision.k 1z p11p 1b p1b k1b k2 2k 2Figure 2.3 Geometry of the A(p, pN) process. The vectors ⃗ b p1 , ⃗ b k1 , <strong>and</strong> ⃗ b k2 are the impact parameters foreach of the three paths for a collision occurr<strong>in</strong>g at ⃗r. z p1 , z k1 , <strong>and</strong> z k2 are the z coord<strong>in</strong>ates of the collisionpo<strong>in</strong>t <strong>in</strong> the respective coord<strong>in</strong>ate systems. θ 1 <strong>and</strong> θ 2 are the angles of the outgo<strong>in</strong>g nucleons relative tothe <strong>in</strong>com<strong>in</strong>g proton direction.It is worth remark<strong>in</strong>g that the eikonal IFSI operators of Eq. (2.55) are one-body operators,i.e., they do not depend on the coord<strong>in</strong>ates (⃗r 2 , ⃗r 3 , . . . , ⃗r A ) of the residual nucleons. The normalizationof the bound-state wave functions simplifies the IFSI factor (2.36) considerably toŜ IFSI (⃗r) = Ŝk1 (⃗r) Ŝk2 (⃗r) Ŝp1 (⃗r) <strong>in</strong> the ROMEA case.


Chapter 2. Relativistic Eikonal A(p, pN) Formalism 21In the relativistic Hartree approximation, real potentials are used to calculate the boundstatewave functions. If the scatter<strong>in</strong>g states are derived us<strong>in</strong>g the same potentials, only elasticrescatter<strong>in</strong>g contributions are taken <strong>in</strong>to account. In general, however, a fraction of the strengthfrom the <strong>in</strong>cident beam is removed from the elastic channel <strong>in</strong>to unobserved <strong>in</strong>elastic ones likethe production of an <strong>in</strong>termediate delta resonance or pion. This local absorption is commonlyimplemented by means of complex or optical potentials obta<strong>in</strong>ed by fitt<strong>in</strong>g nucleon-nucleusscatter<strong>in</strong>g data. In the numerical calculations, we employed the optical potential of van Oers etal. [9] to describe scatter<strong>in</strong>g off 4 He <strong>and</strong> the global S − V parametrizations of Cooper et al. [46]for other target nuclei. Hereafter, the A(p, 2p) calculations which adopt Eq. (2.55) as a start<strong>in</strong>gbasis are labeled the relativistic optical model eikonal approximation (ROMEA).2.3 Relativistic Multiple-Scatter<strong>in</strong>g Glauber ApproximationIn the ROMEA approach, like <strong>in</strong> the DWIA approaches, all the IFSI effects are parametrized <strong>in</strong>terms of mean-field like optical potentials, i.e., the IFSI are seen as a scatter<strong>in</strong>g of the nucleonwith the residual nucleus as a whole. As the energy <strong>in</strong>creases, shorter distances are probed<strong>and</strong> the scatter<strong>in</strong>g with the <strong>in</strong>dividual nucleons becomes more relevant. For proton k<strong>in</strong>eticenergies T p ≥ 1 GeV, the predom<strong>in</strong>antly <strong>in</strong>elastic <strong>and</strong> diffractive character of the underly<strong>in</strong>gelementary proton-nucleon scatter<strong>in</strong>g cross sections makes the Glauber approach [62–64] morenatural. This method reestablishes the l<strong>in</strong>k between proton-nucleus <strong>in</strong>teractions <strong>and</strong> the elementaryproton-proton <strong>and</strong> proton-neutron scatter<strong>in</strong>g. It essentially relies on the eikonal, or,equivalently, the small-angle approximation <strong>and</strong> the assumption of consecutive cumulativescatter<strong>in</strong>gs of a fast nucleon on a composite target conta<strong>in</strong><strong>in</strong>g “frozen” po<strong>in</strong>t scatterers (nucleons).In Glauber theory, the scatter<strong>in</strong>g wave function is related directly to the nucleon-nucleonelastic scatter<strong>in</strong>g data through the <strong>in</strong>troduction of a profile function. This method has a longtradition of successful results <strong>in</strong> high-energy proton-nucleus elastic scatter<strong>in</strong>g (T p > 500 MeV)[68,79], <strong>and</strong> has also been shown to be reliable <strong>in</strong> describ<strong>in</strong>g A(e, e ′ p) processes at high energies[57, 59, 80–88]. The relativistic Glauber formalism exposed here was orig<strong>in</strong>ally developed todescribe A(e, e ′ p) observables [57, 59].2.3.1 Nucleon-Nucleon Scatter<strong>in</strong>gWe start our derivations of a relativistic Glauber formalism by consider<strong>in</strong>g a nucleon-nucleonscatter<strong>in</strong>g process governed by a local Lorentz scalar <strong>and</strong> vector potential V s (r) <strong>and</strong> V v (r). Inthe EA, the scatter<strong>in</strong>g amplitude reads [70]F msm ′ (⃗ ks i , ⃗ k f , E) = − M 〈NΨ (+)〉2π⃗ kf ,m ′ ∣ (βV ∣s + V v ) ∣Φ ⃗ki ,mss, (2.56)with Ψ (+)⃗ kf ,m ′ sthe relativistic scattered state of Eq. (2.51) <strong>and</strong> Φ ⃗ki ,m sthe free Dirac solution (2.8).Note that <strong>in</strong> this section, V s (r) <strong>and</strong> V v (r) represent nucleon-nucleon potentials; whereas <strong>in</strong>


2.3. Relativistic Multiple-Scatter<strong>in</strong>g Glauber Approximation 22Section 2.2, the notation refers to nucleon-nucleus potentials. Some algebraic manipulationslead to the follow<strong>in</strong>g form for the scatter<strong>in</strong>g amplitude [70]〈∫〉F msm ′ (⃗ ks i , ⃗ k f , E) = m ′ iKs ∣ d2π⃗ b e i⃗ ∆·⃗b Γ NN (K, ⃗ b)∣ m s , (2.57)where the profile function is def<strong>in</strong>ed asΓ NN (K, ⃗ b) = 1 − e iχ(K,⃗ b) , (2.58)with the phase-shift function given byχ(K, ⃗ b) = − M NK∫ ∞−∞dz{V c ( ⃗ b, z) + V so ( ⃗ b, z) (⃗σ ·⃗b × K) ⃗ }. (2.59)The Darw<strong>in</strong> term V so ( ⃗ b, z ′ ) (−iKz) present <strong>in</strong> Eq. (2.50) does not contribute to the elastic scatter<strong>in</strong>gamplitude s<strong>in</strong>ce it is an odd function of z.In st<strong>and</strong>ard Glauber theory, knowledge about the NN <strong>in</strong>teraction potentials V c (r) <strong>and</strong>V so (r) is not needed, s<strong>in</strong>ce the phase-shift function χ(K, ⃗ b) can be directly extracted from protonproton<strong>and</strong> proton-neutron scatter<strong>in</strong>g data on the basis of Eq. (2.57). This procedure will beoutl<strong>in</strong>ed <strong>in</strong> what follows.Assum<strong>in</strong>g parity conservation, time-reversal <strong>in</strong>variance, the Pauli pr<strong>in</strong>ciple, <strong>and</strong> isosp<strong>in</strong><strong>in</strong>variance, the most general form for the scatter<strong>in</strong>g amplitude <strong>in</strong> the NN c.m. frame can bewritten <strong>in</strong> terms of five <strong>in</strong>variant amplitudes [68, 89]F ( ⃗ ∆) = A( ⃗ ∆) + B( ⃗ ∆) (⃗σ 1 + ⃗σ 2 ) · ˆn + C( ⃗ ∆) (⃗σ 1 · ˆn) (⃗σ 2 · ˆn)+ D( ⃗ ∆) (⃗σ 1 · ˆm) (⃗σ 2 · ˆm) + E( ⃗ ∆) (⃗σ 1 · ˆl) (⃗σ 2 · ˆl) . (2.60)The <strong>in</strong>itial <strong>and</strong> f<strong>in</strong>al nucleon sp<strong>in</strong> operators are denoted by ⃗σ 1 <strong>and</strong> ⃗σ 2 , <strong>and</strong> ˆn ≡ ⃗ k i × ⃗ k f| ⃗ k i × ⃗ , ˆm ≡k f |⃗ ki − ⃗ k f| ⃗ k i − ⃗ , <strong>and</strong> ˆl ≡ ⃗ k i + ⃗ k fk f | | ⃗ k i + ⃗ . Here, ⃗ k i is the <strong>in</strong>itial, ⃗ kk f | f the f<strong>in</strong>al, <strong>and</strong> ∆ ⃗ the transferred momentum.As such, the NN scatter<strong>in</strong>g amplitude consists of a central term, a sp<strong>in</strong>-orbit term, <strong>and</strong> threeother sp<strong>in</strong>-dependent terms. In pr<strong>in</strong>ciple, the amplitudes A, B, C, D, <strong>and</strong> E can be determ<strong>in</strong>edthrough a complete phase-shift analysis of the NN scatter<strong>in</strong>g data. In practice, most analysesonly <strong>in</strong>clude the central term of the NN amplitude, s<strong>in</strong>ce the small-angle scatter<strong>in</strong>g of protonswith momentum k > 1 GeV/c is assumed to be dom<strong>in</strong>ated by this sp<strong>in</strong>-<strong>in</strong>dependent term.The sp<strong>in</strong>less version of Glauber theory was very successful <strong>in</strong> the analysis of proton-nucleuscross-section data [62, 68, 90]. The surpris<strong>in</strong>g behavior of sp<strong>in</strong> observables <strong>in</strong> NN scatter<strong>in</strong>g[91, 92], however, poses a real challenge. As of today, a quantitative underst<strong>and</strong><strong>in</strong>g of the sp<strong>in</strong>dependence of the NN <strong>in</strong>teraction above 1 GeV does not exist [93]. In this work, all results areobta<strong>in</strong>ed with<strong>in</strong> the framework of sp<strong>in</strong>less Glauber theory.The central term of Eq. (2.60) is parametrized asA( ⃗ ∆) ≡ A( ⃗ ∆ = 0) exp(− β2 NN ∆22). (2.61)


Chapter 2. Relativistic Eikonal A(p, pN) Formalism 23with βNN 2 the slope parameter. This Gaussian parametrization is based on the diffractive pattern,with broad maxima <strong>and</strong> diffractive dips, observed <strong>in</strong> elastic NN collisions at GeV energies.Similar to Fraunhofer diffraction <strong>in</strong> optics, the diffraction phenomenon occurs when thewavelength of the projectile is short compared to the size of the <strong>in</strong>teraction region. The differentialcross section dσ/dt, with t ≡ (k µ f − kµ i )2 the M<strong>and</strong>elstam variable, is extremely forwardpeaked <strong>and</strong> drops exponentially over many orders. The slope parameter βNN 2 describes the tdependence of the elastic NN differential cross section at forward angles:dσ elNNdt≈ dσel NNdt∣ exp ( −βNN 2 |t| ) . (2.62)t=0Us<strong>in</strong>g the optical theorem Im F (θ = 0, φ = 0) = k σtot4π, one f<strong>in</strong>dsA( ∆) ⃗ = k )σtot NN4π(ɛ NN + i) exp(− β2 NN ∆2, (2.63)2with ɛ NN the ratio of the real to the imag<strong>in</strong>ary part of the scatter<strong>in</strong>g amplitude.The <strong>in</strong>verse Fourier transform of Eq. (2.57) br<strong>in</strong>gs about the follow<strong>in</strong>g expression for theprofile function:Γ NN(k, ⃗ )b = σtot NN (k) (1 − iɛ NN (k))4πβNN 2 (k)( )⃗ b2exp −2βNN 2 (k) . (2.64)The profile function for central elastic NN scatter<strong>in</strong>g depends on the momentum k throughthree parameters: the total NN cross sections σNN tot (k), the slope parameters β2 NN(k), <strong>and</strong> theratios of the real to the imag<strong>in</strong>ary part of the scatter<strong>in</strong>g amplitude ɛ NN (k). These parameterscan be determ<strong>in</strong>ed directly from the elementary proton-proton <strong>and</strong> proton-neutron scatter<strong>in</strong>gdata, <strong>and</strong> will be discussed <strong>in</strong> Section 2.3.4.At lower energies, that part of the profile function proportional to ɛ NN (k) is non-Gaussian<strong>and</strong> makes significant contributions to nuclear scatter<strong>in</strong>g. Rather than Eq. (2.64), a parametrization<strong>in</strong> terms of the Arndt NN phases [12,13] is appropriate at lower energies. For the Glaubercalculations presented here, which address higher energies, the Gaussian-like real part ofΓ NN(k, ⃗ )b is the dom<strong>in</strong>ant contributor, <strong>and</strong> the use of Eq. (2.64) is justified.2.3.2 Glauber Multiple-Scatter<strong>in</strong>g Extension of the EAAssum<strong>in</strong>g that the sequential scatter<strong>in</strong>gs are <strong>in</strong>dependent, the eikonal method can easily beextended to multiple scatter<strong>in</strong>g. This constitutes the basis of Glauber theory [62–64]. The scatter<strong>in</strong>goff a composite system (the IFSI <strong>in</strong> A(p, pN) reactions) is modeled as the scatter<strong>in</strong>g ofa fast particle (the <strong>in</strong>cident, scattered, or ejected nucleon) with the scatter<strong>in</strong>g centers of thecomposite system (the spectator nucleons <strong>in</strong> the residual nucleus). We assume that the fastparticle’s momentum is much larger than that of the spectator nucleons. Then, the particle undergoesa negligible deflection <strong>and</strong> its trajectory can be approximated by a straight l<strong>in</strong>e (thisis the EA). In addition, because the time it takes the particle to traverse the nucleus is very


2.3. Relativistic Multiple-Scatter<strong>in</strong>g Glauber Approximation 24small, the spectator nucleons can be approximated by fixed scatter<strong>in</strong>g centers (the so-calledfrozen approximation). As mentioned before, we adopt a sp<strong>in</strong>less Glauber theory, i.e., the fastparticle only <strong>in</strong>teracts with the scatterers by means of two-body sp<strong>in</strong>-<strong>in</strong>dependent <strong>in</strong>teractions.Exchange effects between the fast particle <strong>and</strong> the spectator nucleons are neglected as well.For a multiple-scatter<strong>in</strong>g event lead<strong>in</strong>g from an <strong>in</strong>itial state |i〉 to a f<strong>in</strong>al state |f〉, the Glauberscatter<strong>in</strong>g amplitude readsF multi ( ∆) ⃗ = iK ∫2πwhere the total Glauber phase-shift functionχ tot (K, ⃗ b, ⃗ b 2 , . . . , ⃗ b A ) =d ⃗ b e i⃗ ∆·⃗b 〈f| 1 − e iχtot(K,⃗ b, ⃗ b 2 ,..., ⃗ b A ) |i〉 , (2.65)A∑χ j (K, ⃗ b − ⃗ b j ) , (2.66)j=2is the sum of the phase shifts χ j contributed by each of the spectator scatterers. Here, ⃗ b denotesthe impact parameter of the fast nucleon <strong>and</strong> ( ⃗ b 2 , ⃗ b 3 , . . . , ⃗ b A ) those of the frozen spectator nucleons.The phase-shift additivity property of Eq. (2.66) is a direct result of the follow<strong>in</strong>g assumptions:the one-dimensional nature of the relative motion <strong>and</strong> the neglect of three- <strong>and</strong> morebodyforces, scatterer motion, <strong>and</strong> longitud<strong>in</strong>al momentum transfer. Moreover, as Eq. (2.65) isbased on the eikonal approximation, it is only valid when the energy transfer is small comparedto the <strong>in</strong>cident particle energy, i.e., for elastic <strong>and</strong> mildly <strong>in</strong>elastic collisions. Expression (2.65)does not apply to deeply <strong>in</strong>elastic collisions <strong>in</strong> which the nature of the particles is modified orthe number of particles is altered dur<strong>in</strong>g the collision.Under the assumption of phase-shift additivity, the eikonal wave function of Eq. (2.51) canbe generalized to many-body scatter<strong>in</strong>g. The scatter<strong>in</strong>g wave function for a fast nucleon withmomentum k <strong>and</strong> sp<strong>in</strong> state ∣ 12 m s〉which scatters from A − 1 residual nucleons reads√ []Ψ (+) E + MN 1⃗(⃗r) =k,ms 2M N1E+M N⃗σ · ˆ⃗pŜ e i⃗k·⃗r χ 1 , (2.67)ms2where the operator Ŝ implements the subsequent collisions of the fast nucleon with the frozenspectator nucleonsŜ (⃗r, ⃗r 2 , ⃗r 3 , . . . , ⃗r A ) ≡A∏j=2e −i M NKR z−∞ dz ′ V c( ⃗ b− ⃗ b j ,z ′ −z j ) . (2.68)As <strong>in</strong> the ROMEA framework, only the central sp<strong>in</strong>-<strong>in</strong>dependent contribution V c ( ⃗ b, z) is reta<strong>in</strong>ed,the impulse operator is replaced by the nucleon momentum, <strong>and</strong> the dynamical enhancementof the lower component has been neglected s<strong>in</strong>ce E + M N ≫ |V s (r) − V v (r)| for thehigh energies at which Glauber theory is applied. In this general form, the multiple-scatter<strong>in</strong>gwave function can not be directly related to the <strong>in</strong>dividual profile function for NN scatter<strong>in</strong>g(2.64).


Chapter 2. Relativistic Eikonal A(p, pN) Formalism 25The zero-range approximation along the scatter<strong>in</strong>g direction offers the possibility to expressthe multiple-scatter<strong>in</strong>g wave function <strong>in</strong> terms of the experimentally determ<strong>in</strong>ed profilefunction, thereby avoid<strong>in</strong>g the technical complications with respect to potential scatter<strong>in</strong>g. Thezero-range approximationV c ( ⃗ b − ⃗ b j , z ′ − z j ) ≃ V ⊥c ( ⃗ b − ⃗ b j ) δ(z ′ − z j ) , (2.69)amounts to neglect<strong>in</strong>g the f<strong>in</strong>ite longitud<strong>in</strong>al dimension of the NN <strong>in</strong>teraction region. Afterexp<strong>and</strong><strong>in</strong>g expression (2.68) for Ŝ <strong>and</strong> adopt<strong>in</strong>g the zero-range approximation, one obta<strong>in</strong>sŜ ≃A∏j=2(1 − i M )NK V c ⊥ ( ⃗ b − ⃗ b j ) θ(z ′ − z j ). (2.70)Us<strong>in</strong>g the relation (2.58) between the profile function <strong>and</strong> the phase-shift function, a similarreason<strong>in</strong>g leads toΓ NN (K, ⃗ b) ≃ i M NK V ⊥c ( ⃗ b) . (2.71)F<strong>in</strong>ally, the Glauber operator becomesŜ (⃗r, ⃗r 2 , ⃗r 3 , . . . , ⃗r A ) ≡A∏ [1 − Γ NN(K, ⃗ b − ⃗ )b jj=2]θ (z − z j ), (2.72)where the step function θ (z − z j ) ensures that the nucleon only <strong>in</strong>teracts with other nucleonsif they are localized <strong>in</strong> its forward propagation path.2.3.3 RMSGA for A(p, pN) ReactionsThe Glauber operators <strong>in</strong> Eq. (2.36) take the formsŜ p1 (⃗r, ⃗r 2 , ⃗r 3 , . . . , ⃗r A ) =Ŝ k1 (⃗r, ⃗r 2 , ⃗r 3 , . . . , ⃗r A ) =Ŝ k2 (⃗r, ⃗r 2 , ⃗r 3 , . . . , ⃗r A ) =A∏ [1 − Γ pN(p 1 , ⃗ b − ⃗ )b jj=2A∏j=2A∏j=2]θ (z − z j )( [1 − Γ pN k 1 , ⃗ b ′ − ⃗ )b ′ j θ ( z ′ j − z ′)] ,[ (1 − Γ Nk2 N k 2 , ⃗ b ′′ − ⃗ )b ′′j, (2.73a)θ ( z ′′j − z ′′)] ,(2.73b)(2.73c)where N k2 = p (n) for A(p, 2p) (A(p, pn)) reactions. Further, ⃗r denotes the collision po<strong>in</strong>t<strong>and</strong> (⃗r 2 , ⃗r 3 , . . . , ⃗r A ) are the positions of the frozen spectator protons <strong>and</strong> neutrons <strong>in</strong> the target.The ( ⃗ b, z), ( ⃗ b ′ , z ′ ), <strong>and</strong> ( ⃗ b ′′ , z ′′ ) coord<strong>in</strong>ate systems are def<strong>in</strong>ed as <strong>in</strong> Section 2.2.2. The stepfunctions guarantee that the <strong>in</strong>com<strong>in</strong>g proton can only <strong>in</strong>teract with those spectator nucleonsit encounters before the hard collision <strong>and</strong> that the outgo<strong>in</strong>g protons can only <strong>in</strong>teract with thespectator nucleons they f<strong>in</strong>d <strong>in</strong> their forward propagation paths.


2.3. Relativistic Multiple-Scatter<strong>in</strong>g Glauber Approximation 26Contrary to ROMEA, the Glauber IFSI operators of Eq. (2.73) are genu<strong>in</strong>e A-body operators,so the <strong>in</strong>tegration over the coord<strong>in</strong>ates of the spectator nucleons <strong>in</strong> Eq. (2.36) has to be carriedout explicitly:Ŝ RMSGAIFSI (⃗r) ={A∏ ∫∣d⃗r j φ αj (⃗r j ) ∣ [ (2 1 − Γ pN p 1 , ⃗ b − ⃗ )b jj=2×( [1 − Γ pN k 1 , ⃗ b ′ − ⃗ )b ′ j[ (× 1 − Γ Nk2 N k 2 , ⃗ b ′′ − ⃗ )b ′′jθ ( z ′ j − z ′)]]θ (z − z j )θ ( z ′′j − z ′′)]} . (2.74)This makes the numerical evaluation of the Glauber IFSI factor very challeng<strong>in</strong>g. St<strong>and</strong>ardnumerical <strong>in</strong>tegration techniques were adopted to evaluate the IFSI factor <strong>and</strong> no additionalapproximations, such as the commonly used thickness-function approximation, were <strong>in</strong>troduced.Henceforth, we refer to calculations based on Eq. (2.74) as the relativistic multiple-scatter<strong>in</strong>gGlauber approximation (RMSGA).2.3.4 Glauber ParametersIn contrast to the DWIA <strong>and</strong> ROMEA models, all parameters enter<strong>in</strong>g the RMSGA model canbe obta<strong>in</strong>ed directly from elementary nucleon-nucleon scatter<strong>in</strong>g experiments. When calculat<strong>in</strong>gthe Glauber scatter<strong>in</strong>g wave function for a given momentum k, the follow<strong>in</strong>g <strong>in</strong>put isneeded: the total nucleon-nucleon cross section σNN tot , the slope parameter β2 NN, <strong>and</strong> the ratioof the real to the imag<strong>in</strong>ary part of the scatter<strong>in</strong>g amplitude ɛ NN . We obta<strong>in</strong> the proton-proton<strong>and</strong> proton-neutron parameters through <strong>in</strong>terpolation of the database available from the ParticleData Group [94, 95], while the neutron-neutron scatter<strong>in</strong>g parameters are assumed to beidentical to the proton-proton ones because of isosp<strong>in</strong> symmetry.The measured total (σ tot ) <strong>and</strong> elastic (σ el ) cross sections for proton-proton <strong>and</strong> protonneutronscatter<strong>in</strong>g are shown <strong>in</strong> Fig. 2.4. The nature of the NN <strong>in</strong>teraction changes drastically<strong>in</strong> go<strong>in</strong>g from low to high energies. At proton momenta p ≤ 1 GeV/c, the scatter<strong>in</strong>g of nucleonsis completely elastic; while at higher momenta, the <strong>in</strong>teraction becomes <strong>in</strong>elastic <strong>and</strong>absorptive, as new particles are produced. From 1 GeV/c upward, the total reaction cross sectionrema<strong>in</strong>s almost constant, even though the NN <strong>in</strong>teraction leaves the elastic regime <strong>and</strong>becomes more <strong>and</strong> more <strong>in</strong>elastic.The slope parameters β 2 pp <strong>and</strong> β 2 pn can be found by analyz<strong>in</strong>g the t dependence of the differentialcross sections with the aid of expression (2.62). Here, the sp<strong>in</strong>-dependent terms areassumed to be negligible. For proton k<strong>in</strong>etic energies smaller than 1 GeV, the slope parametersextracted <strong>in</strong> this way differ significantly from the values found directly from experiment<strong>and</strong> phase-shift analysis. This can be attributed to a large contribution of the sp<strong>in</strong>-dependentscatter<strong>in</strong>g amplitude [68]. At higher energies, this difference decreases quickly demonstrat<strong>in</strong>gthat the sp<strong>in</strong> terms are small. Below 1 GeV, values for the slope parameters obta<strong>in</strong>ed through


Chapter 2. Relativistic Eikonal A(p, pN) Formalism 27Cross section (mb)10 210σ totσ elppCross section (mb)10 210σ totσ elpn10 -1 1 10Proton momentum (GeV/c)Figure 2.4 Total <strong>and</strong> elastic cross sections for proton-proton <strong>and</strong> proton-neutron scatter<strong>in</strong>g as a functionof the proton lab momentum. The data were taken from Ref. [95]. The solid (dashed) curve is our globalfit to the elastic (total) cross section.Eq. (2.62) are scarce <strong>and</strong> not free of ambiguities due to sp<strong>in</strong> effects. Therefore, <strong>in</strong> our calculations,the slope parameters are determ<strong>in</strong>ed through the follow<strong>in</strong>g relation( ) 2 ( )σpNtot 1 + ɛ 2 pNβ 2 pN =16π σ elpN. (2.75)This parametrization can be derived from the theoretical shape of the elastic pN cross sectionas follows. Expression (2.63) for the elastic scatter<strong>in</strong>g amplitude leads to( ) 2 ( )dσpNeld(∆ 2 ) = π ∣ ∣ ∣∣A( ∆) ⃗ ∣∣2 σpNtot 1 + ɛ 2 pN=k 2 exp ( −β 216πpN∆ 2) . (2.76)Integrat<strong>in</strong>g this st<strong>and</strong>ard high-energy approximation of the elastic differential cross sectionresults <strong>in</strong>∫σpN el =(dσelpNd(∆ 2 ) d(∆2 ) =σ totpN) 2 ( )1 + ɛ 2 pN16πβ 2 pN, (2.77)so that the slope parameter is given by Eq. (2.75). In Fig. 2.5, the slope parameters obta<strong>in</strong>edthrough this expression are compared with those determ<strong>in</strong>ed directly through Eq. (2.62).


2.4. Approximated RMSGA 2812121010β 2 pp [(GeV/c) -2 ]864β 2 pn [(GeV/c) -2 ]8642201Proton momentum (GeV/c)01Proton momentum (GeV/c)Figure 2.5 The slope parameters βpp 2 <strong>and</strong> βpn. 2 The curves are obta<strong>in</strong>ed with Eq. (2.75) <strong>and</strong> the global fitsto σpN tot,σel pN , <strong>and</strong> ɛ pN shown <strong>in</strong> Figs. 2.4 <strong>and</strong> 2.6. The data are from Refs. [96] (proton-proton) <strong>and</strong> [97](proton-neutron) <strong>and</strong> were determ<strong>in</strong>ed through the small-angle t dependence (2.62) of the measureddifferential cross sections.Fig. 2.6 displays the ratios of the real to the imag<strong>in</strong>ary part of the scatter<strong>in</strong>g amplitude.In our formalism, global fits to the data are used. The ɛ pN parameters can also be evaluatedfrom dispersion relations, which provide a connection between the real <strong>and</strong> imag<strong>in</strong>ary partof the scatter<strong>in</strong>g amplitude, <strong>in</strong> comb<strong>in</strong>ation with the optical theorem. The dispersion-relationpredictions for ɛ pp <strong>and</strong> ɛ pn [98, 99] are <strong>in</strong> good agreement with the experimental values.2.4 Approximated RMSGAIn this section, the RMSGA IFSI factor of Eq. (2.74) is rewritten <strong>in</strong> a numerically more convenientform by adopt<strong>in</strong>g some approximations. First, the squared wave functions ∣ ∣φ αj (⃗r j ) ∣ 2 ofthe spectator protons (neutrons) are replaced by ρ p (⃗r j ) /Z res (ρ n (⃗r j ) /N res ). Here, ρ p <strong>and</strong> ρ n arethe proton <strong>and</strong> neutron density of the residual nucleus. Z res <strong>and</strong> N res are the number of protons


Chapter 2. Relativistic Eikonal A(p, pN) Formalism 29ε pp21.510.50-0.5-1-1.51.510.5ε pn0-0.5-1-1.510 -1 1Proton momentum (GeV/c)Figure 2.6 The ratios of the real to the imag<strong>in</strong>ary part of the central scatter<strong>in</strong>g amplitude A(∆) forproton-proton <strong>and</strong> proton-neutron scatter<strong>in</strong>g as a function of the proton lab momentum. The curves areglobal fits to the data. The data are from Ref. [68].<strong>and</strong> neutrons <strong>in</strong> the residual nucleus. This leaves us withŜIFSIRMSGA×(⃗r) ≃{ ∫ρ p (⃗r j )[ (d⃗r j 1 − Γ pp p 1 ,Z ⃗ b − ⃗ )b j θ (z − z j )res[ (× 1 − Γ Nk2 p k 2 , ⃗ b ′′ − ⃗ )b ′′jθ ( z ′′j − z ′′)]} Z res] ( [1 − Γ pp k 1 , ⃗ b ′ − ⃗ )b ′ j θ ( z ′ j − z ′)]{ ∫ρ n (⃗r j )[ (d⃗r j 1 − Γ pn p 1 ,N ⃗ b − ⃗ ) ] (b j θ (z − z j )[1 − Γ pn k 1 , ⃗ b ′ − ⃗ )b ′ j θ ( z ′ j − z ′)]res[ (× 1 − Γ Nk2 n k 2 , ⃗ b ′′ − ⃗ )b ′′jθ ( z ′′j − z ′′)]} N res. (2.78)


2.5. Second-Order Eikonal Corrections 30Second, one(assumes that the proton <strong>and</strong> neutron densities are slowly vary<strong>in</strong>g functions of ⃗ b,while Γ NN k, ⃗ b − ⃗ )b j is sharply peaked at ⃗ b − ⃗ b j = ⃗0. Under this assumption, one can makethe follow<strong>in</strong>g approximation∫(d⃗r j ρ N (⃗r j ) Γ pN p 1 , ⃗ b − ⃗ )b j θ (z − z j )= σtot pN (p 1) (1 − iɛ pN (p 1 ))4π β 2 pN (p 1)≃1 2 σtot pN (p 1 ) (1 − iɛ pN (p 1 ))∫ z−∞∫ z−∞dz j ρ N(⃗b, zj) ∫ d ⃗ b j expdz j ρ N(⃗b, zj)(− (⃗ b − ⃗ )b j ) 22βpN 2 (p 1), (2.79)<strong>and</strong> a similar approximation for the terms l<strong>in</strong>ear <strong>in</strong> Γ pN(k 1 , ⃗ b ′ − ⃗ b ′ j <strong>and</strong> Γ Nk2 Nwhile higher-order terms are neglected. Insert<strong>in</strong>g this <strong>in</strong> Eq. (2.78) yieldsŜIFSI RMSGA′ (⃗r) ={1 − σtot pp (p 1 ) (1 − iɛ pp (p 1 ))2Z res∫ z−∞∫− σtot pp (k 1 ) (1 − iɛ pp (k 1 )) +∞2Z resz ′dz j ρ p(⃗b, zj)dz ′ j ρ p(⃗b ′ , z ′ j− σtot N k2 p (k ∫2) (1 − iɛ Nk2 p (k 2 )) +∞ ( ) } Z resdz ′′j ρ p ⃗b ′′ , z ′′j2Z res z ′′))(k 2 , ⃗ b ′′ − ⃗ )b ′′j ,× factor for the spectator neutrons . (2.80)In the more frequently adopted exponential form, this reads [55, 100]Ŝ RMSGA′IFSI (⃗r) = ∏N=p,ne − 1 2 σtot pN (p 1) (1−iɛ pN (p 1 )) R z−∞ dz j ρ N ( ⃗ b,z j )× e − 1 2 σtot pN (k 1) (1−iɛ pN (k 1 )) R +∞z ′ dz ′ j ρ N ( ⃗ b ′ ,z ′ j )× e − 1 2 σtot N k2 N (k 2) (1−iɛ Nk2 N (k 2 )) R +∞z ′′ dz ′′j ρ N ( ⃗ b ′′ ,z ′′j ) . (2.81)The IFSI operator of Eq. (2.74) can thus be reduced to a one-body operator. Henceforth, calculationsbased on Eq. (2.81) are labeled as RMSGA ′ .2.5 Second-Order Eikonal CorrectionsBecause of its numerous advantages, the eikonal approximation has a long history of successfulresults <strong>in</strong> describ<strong>in</strong>g A(p, pN) reactions as well as other scatter<strong>in</strong>g processes like heavy-ion collisions,photo- <strong>and</strong> electro-<strong>in</strong>duced nucleon-knockout reactions. The eikonal scatter<strong>in</strong>g wavefunctions are derived by l<strong>in</strong>eariz<strong>in</strong>g the cont<strong>in</strong>uum wave equation for the <strong>in</strong>teract<strong>in</strong>g particle.Hence, the solution is only valid to first order <strong>in</strong> 1/k, with k the particle’s momentum, <strong>and</strong>the EA is suited for the description of high-energy scatter<strong>in</strong>g. To extend the applicability tolower energies, Wallace [101–105] has developed systematic corrections to the eikonal scatter<strong>in</strong>gamplitude. Several authors have <strong>in</strong>vestigated the effect of higher-order eikonal corrections


Chapter 2. Relativistic Eikonal A(p, pN) Formalism 31<strong>in</strong> elastic nuclear scatter<strong>in</strong>g by protons, antiprotons, <strong>and</strong> α particles [106, 107], heavy-ion scatter<strong>in</strong>g[108–111], <strong>and</strong> <strong>in</strong>clusive electron-nucleus scatter<strong>in</strong>g [112]. In this section, we develop asecond-order correction to the ROMEA framework of Section 2.2. Our formalism builds uponthe work of Baker [113], where an eikonal approximation for potential scatter<strong>in</strong>g was derivedto second order <strong>in</strong> 1/k. This approach is extended to account for the effect of the sp<strong>in</strong>-orbitpotential.Like <strong>in</strong> the ROMEA approach of Section 2.2, the start<strong>in</strong>g po<strong>in</strong>t is the Schröd<strong>in</strong>ger-likeequation (2.42) for the upper component u (+)⃗ k,ms(⃗r). In the sp<strong>in</strong>-orbit (V so (r) ⃗σ · ⃗L) <strong>and</strong> Darw<strong>in</strong>(V so (r) (−i⃗r · ˆ⃗p)) terms, as well as <strong>in</strong> the lower component (2.43), the momentum operator ˆ⃗p isreplaced by the asymptotic momentum ⃗ k, i.e., the EMA is adopted. For the upper component,one postulates a solution of the formu (+)⃗(⃗r) ≡ N η(⃗r) e i⃗k·⃗r χ 1 , (2.82)k,ms ms2i.e., a plane wave modulated by an eikonal factor η(⃗r). Here, N is a normalization factor.In the ROMEA approach, which adopts the first-order eikonal approximation, the Schröd<strong>in</strong>ger-typeequation (2.42) was then l<strong>in</strong>earized <strong>in</strong> ˆ⃗p lead<strong>in</strong>g to the solution (2.50). Despite the factthat it is reta<strong>in</strong>ed as an exponential phase, this solution is, strictly speak<strong>in</strong>g, only valid up tofirst order <strong>in</strong> V opt /k, with V opt ( ⃗ b, z) = V c ( ⃗ b, z) + V so ( ⃗ b, z) (⃗σ · ⃗b × ⃗ k − ikz). Mathematically,the exponential expression is not justified. However, it is commonly used because physical<strong>in</strong>tuition dictates that the effect of the scatter<strong>in</strong>g is to modulate the <strong>in</strong>com<strong>in</strong>g plane wave by aphase change.In what follows, we will derive an expression for the eikonal factor η(⃗r) that is valid upto order V opt /k 2 . The momentum dependence <strong>in</strong> the sp<strong>in</strong>-orbit <strong>and</strong> Darw<strong>in</strong> terms makes thatthese terms are reta<strong>in</strong>ed up to order V so /k, while central terms are <strong>in</strong>cluded up to order V c /k 2 .Note that the expansion is not expressed <strong>in</strong> terms of the Lorentz scalar <strong>and</strong> vector potentials V s<strong>and</strong> V v . Look<strong>in</strong>g for a solution of the form (2.82) for the Schröd<strong>in</strong>ger-like equation (2.42), Bakerarrived at the follow<strong>in</strong>g equation for the eikonal factor (see Eq. (14) of Ref. [113]):∫ zη( ⃗ b, z) = 1 − i dz ′ V opt (v⃗ b, z ′ ) η( ⃗ b, z ′ ) + 1−∞2kv V opt( ⃗ b, z) η( ⃗ b, z)+ 1 ∫ z( 1dz ′ (z − z ′ )2kv −∞b + ∂ ) ∂(V opt (∂b ∂b⃗ b, z ′ ) η( ⃗ )b, z ′ ) , (2.83)where v = k/M N . Note that, apart from dropp<strong>in</strong>g contributions of order V opt /k 3 <strong>and</strong> higher,no additional assumptions were made when deriv<strong>in</strong>g Eq. (2.83). In Ref. [113], Eq. (2.83) wassubsequently solved for spherically symmetric potentials. The sp<strong>in</strong>-orbit <strong>and</strong> Darw<strong>in</strong> terms,however, break the spherical symmetry <strong>and</strong> a novel method to solve Eq. (2.83) is required.To that purpose, we assume that the derivative of the function η is of higher order <strong>in</strong> 1/kthan η itself (as is true for the ROMEA solution (2.50)). This allows us to drop the ∂η/∂b contri-


2.5. Second-Order Eikonal Corrections 32bution <strong>in</strong> the last term of the right-h<strong>and</strong> side of Eq. (2.83) s<strong>in</strong>ce it is of order V opt /k 3 or higher:∫1 z( 1dz ′ (z − z ′ )2kv −∞b + ∂ ) ∂(V opt (∂b ∂b⃗ b, z ′ ) η( ⃗ )b, z ′ )(1 1=2kv b + ∂ ) ∫ zdz ′ (z − z ′ )∂b −∞[∂(× V c (∂b⃗ b, z ′ ) + V so ( ⃗ b, z ′ ) (⃗σ ·⃗b × ⃗ k − ikz )) ]′ η( ⃗ b, z ′ ) . (2.84)Spherical symmetry implies that z ′ ∂V c ( ⃗ b, z ′ )/∂b = b ∂V c ( ⃗ b, z ′ )/∂z ′ . Hence, the z ′ ∂V c /∂b term<strong>in</strong> Eq. (2.84) can be written as− 1 ( 12kv b + ∂ ) ∫ zdz ′ b ∂V c( ⃗ b, z ′ )∂b −∞ ∂z ′ η( ⃗ b, z ′ )= − 1 ( 12kv b + ∂ ) ∫ zdz ′ b ∂ (∂b −∞ ∂z ′ V c ( ⃗ b, z ′ ) η( ⃗ )b, z ′ )[ (= − 1 2 + b ∂ ) ]V c (2kv ∂b⃗ b, z) η( ⃗ b, z) . (2.85)In the first step, the fact that the derivative ∂η/∂z ′ is of higher order was used to turn the<strong>in</strong>tegr<strong>and</strong> <strong>in</strong>to an exact differential. A similar reason<strong>in</strong>g, followed by <strong>in</strong>tegration by parts,leads to(1 12kv b + ∂ ) ∫ zdz ′ (z − z ′ ) ∂V so( ⃗ b, z ′ )(−ikz ′ ) η(∂b −∞∂b⃗ b, z ′ )= − i ∫ [z(dz ′ 2 + b ∂ ) ]V so (2v −∞ ∂b⃗ b, z) η( ⃗ b, z ′ ) , (2.86)for the Darw<strong>in</strong> term of Eq. (2.84). As a result, Eq. (2.83) takes the formη( ⃗ b, z) =1 − i ∫ [z(dz ′ V opt (v⃗ b, z ′ ) η( ⃗ b, z ′ ) − 1 1 + b ∂ ) ]V c (−∞2kv ∂b⃗ b, z) η( ⃗ b, z)+ z (1 + b ∂ ) ∫ zdz ′ ∂V c( ⃗ b, z ′ )η(2kvb ∂b −∞ ∂b⃗ b, z ′ )+ 12kv V so( ⃗ b, z) (⃗σ ·⃗b × ⃗ k − ikz) η( ⃗ b, z)+ 1 (1 + b ∂ ) ∫ [zdz ′ (z − z ′ ∂() V so (2kvb ∂b −∞∂b⃗ b, z ′ ) ⃗σ ·⃗b × ⃗ k) ] η( ⃗ b, z ′ )− i ∫ [z(dz ′ 2 + b ∂ ) ]V so (2v −∞ ∂b⃗ b, z) η( ⃗ b, z ′ ) . (2.87)We look for a solution of the form(η( ⃗ b, z) = f( ⃗ b, z) exp − i ∫ z)d¯z V opt (v⃗ b, ¯z) f( ⃗ b, ¯z)−∞= f( ⃗ (b, z) exp i S( ⃗ )b, z) , (2.88)


Chapter 2. Relativistic Eikonal A(p, pN) Formalism 33which reduces to the ROMEA result of Eq. (2.50) when terms of higher order than V opt /k areneglected. Accord<strong>in</strong>gly, the function f( ⃗ b, z) should be of the form f = 1 + O(V opt /k 2 ). Substitut<strong>in</strong>g(2.88) <strong>in</strong>to Eq. (2.87) <strong>and</strong> multiply<strong>in</strong>g by e −i S(⃗b,z) on the right yields[ (f( ⃗ b, z) = 1 − 1 1 + b ∂ ) ]V c (2kv ∂b⃗ b, z) f( ⃗ b, z)+ z (1 + b ∂ ) ∫ zdz ′ ∂V c( ⃗ b, z ′ )f(2kvb ∂b −∞ ∂b⃗ b, z ′ )+ 12kv V so( ⃗ b, z) (⃗σ ·⃗b × ⃗ k − ikz) f( ⃗ b, z)+ 1 (1 + b ∂ ) ∫ [zdz ′ (z − z ′ ∂() V so (2kvb ∂b −∞∂b⃗ b, z ′ ) ⃗σ ·⃗b × ⃗ k) ] f( ⃗ b, z ′ )− i ∫ [z(dz ′ 2 + b ∂ ) ]V so (2v −∞ ∂b⃗ b, z) f( ⃗ b, z ′ ) . (2.89)In deriv<strong>in</strong>g this equation, we set e i S(⃗ b,z ′) e −i S(⃗ b,z) equal to 1, s<strong>in</strong>ce higher-order terms are neglected.The difficulty <strong>in</strong> solv<strong>in</strong>g for f( ⃗ b, z) is that Eq. (2.89) is an <strong>in</strong>tegral equation. An expressionfor f( ⃗ b, z) can, however, be readily obta<strong>in</strong>ed by add<strong>in</strong>g (1−f) terms, which <strong>in</strong>troduce onlyhigher-order terms, to the right-h<strong>and</strong> side of Eq. (2.89). This is permitted s<strong>in</strong>ce the solution isonly determ<strong>in</strong>ed up to order V opt /k 2 . With this manipulation, the function f becomesf( ⃗ b, z) = 1 − 1 (1 + b ∂ )V c (2kv ∂b⃗ b, z) +z (1 + b ∂ ) ∫ zdz ′ ∂V c( ⃗ b, z ′ )2kvb ∂b −∞ ∂b+ 12kv V so( ⃗ b, z) (⃗σ ·⃗b × ⃗ k − ikz)+ 1 (1 + b ∂ ) ∫ zdz ′ (z − z ′ ) ∂ (V so (2kvb ∂b −∞ ∂b⃗ b, z ′ ) ⃗σ ·⃗b × ⃗ )k− i ∫ z(dz ′ 2 + b ∂ )V so (2v∂b⃗ b, z) . (2.90)−∞The eikonal factor of Eq. (2.88) whereby f is determ<strong>in</strong>ed by (2.90) is a solution of the <strong>in</strong>tegralequation (2.83) to order V opt /k 2 <strong>and</strong> <strong>in</strong>deed reduces to the ROMEA result (2.50) when truncatedat order V opt /k. Furthermore, it can be easily verified that the derivative of η is of higher order<strong>in</strong> V opt /k than η itself. Henceforth, results obta<strong>in</strong>ed with the eikonal factor given by Eqs. (2.88)<strong>and</strong> (2.90) are dubbed as the second-order relativistic optical model eikonal approximation(SOROMEA).


2.5. Second-Order Eikonal Corrections 34


Chapter 3Initial- <strong>and</strong> F<strong>in</strong>al-State Interactions <strong>and</strong>A(p, pN) Differential Cross SectionsIn the previous chapter, we have developed a cross-section factorized framework for describ<strong>in</strong>gA(p, pN) processes <strong>and</strong> have discussed various methods to treat the IFSI. In this chapter,we will put our framework to a str<strong>in</strong>gent test by compar<strong>in</strong>g our calculations with exclusiveA(p, pN) cross-section data that have been collected at various facilities. The ma<strong>in</strong> focus is onthe optical-potential approach, as this method turns out to be the most suitable for the descriptionof the IFSI for the k<strong>in</strong>ematical sett<strong>in</strong>gs of the experiments discussed <strong>in</strong> this chapter.Section 3.1 is devoted to a presentation of optical-potential <strong>and</strong> Glauber results for the IFSIfactor. This is a function which accounts for all IFSI effects when comput<strong>in</strong>g the A(p, pN) observables.In Section 3.2, the radial <strong>and</strong> polar-angle contributions to the A(p, pN) cross sectionsare studied. F<strong>in</strong>ally, <strong>in</strong> Section 3.3, the optical-potential predictions of our model are comparedwith cross-section data that have been collected at the Petersburg <strong>Nuclear</strong> <strong>Physics</strong> Institute(PNPI), the Enrico Fermi Institute (EFI), <strong>and</strong> the Tri-University Meson Facility (TRIUMF). First,we present our calculations for the 12 C, 16 O, <strong>and</strong> 40 Ca(p, 2p) <strong>and</strong> (p, pn) PNPI data for 1 GeV<strong>in</strong>com<strong>in</strong>g proton energies [114]. Second, the results from the 40 Ca(p, 2p) experiment performedat EFI [115], one of the first of its k<strong>in</strong>d, are scrut<strong>in</strong>ized. F<strong>in</strong>ally, our predictions for scatter<strong>in</strong>goff 16 O <strong>and</strong> 4 He are compared to the TRIUMF data of McDonald [116], Kitch<strong>in</strong>g [8], <strong>and</strong> vanOers [9].Most results presented <strong>in</strong> this chapter were orig<strong>in</strong>ally published <strong>in</strong> [117].3.1 Numerical Results for the IFSI FactorIn this section, results for the IFSI factor (2.36) are given for the knockout of nucleons from theFermi level <strong>in</strong> 12 C, 16 O, <strong>and</strong> 40 Ca, at an <strong>in</strong>cident energy T p1 = 1 GeV <strong>and</strong> a scattered proton k<strong>in</strong>eticenergy T k1 = 870 MeV. Thereby, we adopt coplanar scatter<strong>in</strong>g angles (θ 1 , θ 2 ) = (13.4 ◦ , 67 ◦ )on opposite sides of the <strong>in</strong>cident beam, i.e., k<strong>in</strong>ematics correspond<strong>in</strong>g with the PNPI experimentof Ref. [114]. All IFSI effects are <strong>in</strong>cluded <strong>in</strong> the IFSI factor S IFSI (⃗r). Note that <strong>in</strong> the ab-35


3.1. The IFSI Factor 36sence of <strong>in</strong>itial- <strong>and</strong> f<strong>in</strong>al-state <strong>in</strong>teractions the real part of the IFSI factor equals one, whereasthe imag<strong>in</strong>ary part vanishes identically.The A(p, pN) IFSI factor is a function of three <strong>in</strong>dependent variables (r, θ, φ). The z axis ischosen along the direction of the <strong>in</strong>com<strong>in</strong>g beam ⃗p 1 , the y axis lies along ⃗p 1 × ⃗ k 1 , <strong>and</strong> the xaxis lies <strong>in</strong> the scatter<strong>in</strong>g plane def<strong>in</strong>ed by the proton momenta ⃗p 1 <strong>and</strong> ⃗ k 1 . θ <strong>and</strong> φ denote thepolar <strong>and</strong> azimuthal angles with respect to the z axis <strong>and</strong> the x axis, respectively. The radialcoord<strong>in</strong>ate r represents the distance relative to the center of the target nucleus.3.1.1 Polar-Angle DependenceTo ga<strong>in</strong> a better <strong>in</strong>sight <strong>in</strong>to the dependence of the IFSI factor on r, θ, <strong>and</strong> φ, we calculated thecontribution of the three distort<strong>in</strong>g functions Ŝp1, Ŝ k1 , <strong>and</strong> Ŝk2 to the IFSI factor. In Figs. 3.1<strong>and</strong> 3.2 results are displayed for the computed real <strong>and</strong> imag<strong>in</strong>ary part of S IFSI (r, θ, φ = 0) forproton emission from the Fermi level <strong>in</strong> 12 C. The results were computed with<strong>in</strong> the ROMEAframework, us<strong>in</strong>g the EDAD1 optical-potential parametrization of [46].The θ dependence can be <strong>in</strong>terpreted as follows. For a given r, the distance that the <strong>in</strong>com<strong>in</strong>gproton travels through the target nucleus before collid<strong>in</strong>g hard with a target nucleondecreases with <strong>in</strong>creas<strong>in</strong>g angle θ. As a consequence, small values of θ <strong>in</strong>duce the largest ISI.For the FSI of the scattered proton, the opposite holds true, <strong>and</strong> θ = 180 ◦ for a large r valuecorresponds to an event whereby the hard collision transpires at the outskirts of the nucleus<strong>and</strong> the scattered proton has to travel through the whole nucleus before it becomes asymptoticallyfree, thus giv<strong>in</strong>g rise to the smallest (largest) values for the real (imag<strong>in</strong>ary) part of theIFSI factor. Unlike the scattered proton, which moves almost coll<strong>in</strong>ear to the z axis, the ejectednucleon leaves the nucleus under a large scatter<strong>in</strong>g angle θ 2 . Hence, the FSI are m<strong>in</strong>imal forθ close to 0 ◦ or 180 ◦ <strong>and</strong> maximal for θ around 180 ◦ − θ 2 . F<strong>in</strong>ally, the θ dependence of thecomplete IFSI factor is the result of the <strong>in</strong>terplay between the three distort<strong>in</strong>g effects, with thestrongest scatter<strong>in</strong>g <strong>and</strong> absorption observed at θ close to 0 ◦ , 180 ◦ − θ 2 , <strong>and</strong> 180 ◦ .3.1.2 Radial DependenceFig. 3.3 displays the real part of the 40 Ca IFSI factor as a function of r at various values of θ.The ROMEA calculations were performed for the same k<strong>in</strong>ematics as <strong>in</strong> Figs. 3.1 <strong>and</strong> 3.2, <strong>and</strong>employed the EDAI optical-potential fit of [46].The upper left panel suggest that the ISI effects <strong>in</strong>crease with grow<strong>in</strong>g r for θ = 0 ◦ . Forθ = 45 ◦ <strong>and</strong> <strong>in</strong>creas<strong>in</strong>g r, <strong>in</strong>itially, the grow<strong>in</strong>g distance the proton has to travel through thenucleus leads to a decrease of the real part of Ŝp1. This is followed by an <strong>in</strong>crease for larger rup to Ŝp1 = 1. This reduction <strong>in</strong> ISI effects with <strong>in</strong>creas<strong>in</strong>g r is brought about by the <strong>in</strong>com<strong>in</strong>gproton’s path through the nucleus mov<strong>in</strong>g away from the nuclear <strong>in</strong>terior <strong>and</strong> closer to theless dense nuclear surface. The other curves of the upper left figure reveal a general trend for90 ◦ ≤ θ ≤ 180 ◦ : as r <strong>in</strong>creases, the real part of the IFSI factor grows correspond<strong>in</strong>gly. As canbe appreciated from Fig. 3.3 as well as from the previous figures, the global behavior of the Ŝk1


Chapter 3. IFSI <strong>and</strong> A(p, pN) Differential Cross Sections 37real part of S IFSI10.80.60.40.2real part of S IFSI10.80.60.40.2150θ (deg)10050012345r (fm)6150θ (deg)10050012345r (fm)6real part of S IFSI0.80.60.4real part of S IFSI0.80.60.40.2150θ (deg)10050012345r (fm)6150θ (deg)10050012345r (fm)6Figure 3.1 The radial <strong>and</strong> polar-angle dependence of the real part of the IFSI factor S IFSI <strong>in</strong> the scatter<strong>in</strong>gplane (φ = 0 ◦ ) for proton knockout from the Fermi level <strong>in</strong> 12 C. The upper left panel is the contributionfrom the imp<strong>in</strong>g<strong>in</strong>g proton (Ŝp1), while the upper right panel shows the effect of the FSI of the scatteredproton (Ŝk1). In the bottom left figure, the effect of the FSI of the ejected proton (Ŝk2) is presented <strong>and</strong>the bottom right figure shows the complete IFSI factor (Ŝk1 Ŝk2 Ŝp1). The k<strong>in</strong>ematics was T p1 = 1 GeV,T k1 = 870 MeV, θ 1 = 13.4 ◦ , <strong>and</strong> θ 2 = 67 ◦ .factor describ<strong>in</strong>g the scattered proton’s FSI can be related to that of the ISI factor Ŝp1 through thesubstitution θ → 180 ◦ − θ. This approximate symmetry can be attributed to the small scatter<strong>in</strong>gangle θ 1 , i.e., the scattered proton leaves the nucleus almost parallel to the <strong>in</strong>com<strong>in</strong>g proton’sdirection. In the bottom left panel, the additional curve (θ = 115 ◦ , i.e., close to 180 ◦ − θ 2 )represents the situation of maximal FSI of the ejected nucleon. For this θ value, the path of theejected nucleon passes through the center of the nucleus, <strong>and</strong> the distance traveled throughthe nucleus <strong>in</strong>creases with r. Accord<strong>in</strong>gly, the real part of Ŝk2 is a monotonously decreas<strong>in</strong>gfunction of r. The other extreme is the θ = 0 ◦ case, where <strong>in</strong>creas<strong>in</strong>g r means less FSI. For theother θ values, the absorption reaches its maximum for some <strong>in</strong>termediate r value. Aga<strong>in</strong>, thecomb<strong>in</strong>ation of Ŝp1, Ŝk1, <strong>and</strong> Ŝk2 determ<strong>in</strong>es the total IFSI factor, with the strongest attenuationpredicted <strong>in</strong> the nuclear <strong>in</strong>terior.


3.1. The IFSI Factor 38imag<strong>in</strong>ary part of S IFSI0-0.05-0.1-0.15-0.2150100θ (deg)50012345r (fm)6imag<strong>in</strong>ary part of S IFSI0-0.05-0.1-0.15-0.2150θ (deg)10050012345r (fm)6imag<strong>in</strong>ary part of S IFSI0.30.20.1150100θ (deg)50012345r (fm)6imag<strong>in</strong>ary part of S IFSI0.30.20.10-0.1-0.2150θ (deg)10050012345r (fm)6Figure 3.2 As <strong>in</strong> Fig. 3.1, but for the imag<strong>in</strong>ary part of the IFSI factor.3.1.3 Azimuthal-Angle DependenceThe dependence of the IFSI factor on the azimuthal angle of the collision po<strong>in</strong>t is quite straightforward.One representative result is displayed <strong>in</strong> Fig. 3.4. Here, cos φ ≥ 0 (cos φ ≤ 0) refers toa situation where the hard NN collision occurs <strong>in</strong> the upper (lower) hemisphere with respectto the yz plane (see Fig. 2.3 for a collision po<strong>in</strong>t located <strong>in</strong> the upper hemisphere). Because ofthe cyl<strong>in</strong>drical symmetry about the z axis, the factor describ<strong>in</strong>g the ISI of the <strong>in</strong>com<strong>in</strong>g protonis <strong>in</strong>dependent of φ. Regard<strong>in</strong>g the scattered proton, we observe the least FSI <strong>in</strong> the upperhemisphere, s<strong>in</strong>ce the proton then avoids pass<strong>in</strong>g through the highly absorb<strong>in</strong>g nuclear <strong>in</strong>terior.For the ejected nucleon, the opposite applies, <strong>and</strong> the strongest FSI effects are found forφ = 0 ◦ . As the xz plane is def<strong>in</strong>ed as the scatter<strong>in</strong>g plane, the IFSI factor possesses the symmetryS IFSI (r, θ, 2π − φ) = S IFSI (r, θ, φ) for coplanar scatter<strong>in</strong>g.3.1.4 Level <strong>and</strong> A DependenceResults for the emission from levels other than the Fermi level have not been plotted here,because it turns out that the IFSI factors hardly depend on the s<strong>in</strong>gle-particle level <strong>in</strong> which thestruck nucleon resides. Also, the IFSI factors for neutron emission are almost identical to the


Chapter 3. IFSI <strong>and</strong> A(p, pN) Differential Cross Sections 39110.90.90.80.8real part of S IFSI0.70.60.50.40.3real part of S IFSI0.70.60.50.40.30.20.20.10.101 2 3 4 5 6 7 8r (fm)01 2 3 4 5 6 7 8r (fm)real part of S IFSI10.90.80.70.60.50.40.3θ = 0 oθ = 45 oθ = 90 oθ = 135 oθ = 180 oθ = 115 oreal part of S IFSI10.90.80.70.60.50.40.30.20.20.10.101 2 3 4 5 6 7 8r (fm)01 2 3 4 5 6 7 8r (fm)Figure 3.3 The radial dependence of the real part of the IFSI factor S IFSI <strong>in</strong> the scatter<strong>in</strong>g plane (φ = 0 ◦ )for neutron knockout from the Fermi level <strong>in</strong> 40 Ca. The upper left, upper right, <strong>and</strong> bottom left panelsdisplay Ŝp1, Ŝ k1 , <strong>and</strong> Ŝk2, respectively; while the complete IFSI factor is shown <strong>in</strong> the bottom rightpicture.correspond<strong>in</strong>g IFSI factors for proton knockout <strong>and</strong>, as expected, the overall effect of IFSI ismore pronounced for heavier target nuclei due to the larger number of scatter<strong>in</strong>g centers. Thelatter is illustrated <strong>in</strong> Fig. 3.5, which shows the ratio of the 40 Ca absolute value of the IFSI factorto the 12 C one.3.1.5 Comparison between ROMEA <strong>and</strong> RMSGA CalculationsIn this subsection, we <strong>in</strong>vestigate the sensitivity of the computed IFSI factors to the adoptedparametrizations for the optical potentials <strong>and</strong> compare the ROMEA results with the RMSGApredictions. As can be seen <strong>in</strong> Fig. 3.6, the IFSI factor depends on whether A-dependent(EDAD1/EDAD2) or A-<strong>in</strong>dependent (EDAI) fits for the potentials are selected, but the globalfeatures are comparable. Fig. 3.7, as contrasted to Fig. 3.1, demonstrates that the RMSGAmethod adequately describes the ISI of the <strong>in</strong>com<strong>in</strong>g proton <strong>and</strong> the FSI of the scattered proton.However, the discrepancies between ROMEA <strong>and</strong> RMSGA become significant <strong>in</strong> the calculationof the FSI of the ejected nucleon (note the different scales <strong>in</strong> the bottom left panels ofFigs. 3.1 <strong>and</strong> 3.7), <strong>and</strong>, therefore, also <strong>in</strong> the complete IFSI factor. The noted difference is at-


3.2. Radial <strong>and</strong> Polar-Angle Contributions to A(p, pN) Cross Sections 40real part of S IFSI0.80.60.40.2real part of S IFSI0.80.60.40.2300φ (deg)2001000030θ (deg)6090120150180300φ (deg)2001000030θ (deg)6090120150180real part of S IFSI0.90.80.70.60.50.4real part of S IFSI0.60.50.40.30.20.1300φ (deg)2001000030θ (deg)6090120150180300φ (deg)2001000030θ (deg)6090120150180Figure 3.4 The polar- <strong>and</strong> azimuthal-angle dependence of the real part of S IFSI (r = 3 fm, θ, φ) for protonknockout from the Fermi level <strong>in</strong> 16 O. K<strong>in</strong>ematics as <strong>in</strong> Fig. 3.1. ROMEA calculation with the EDAD2optical potential [46]. As <strong>in</strong> the previous figures, the upper left, upper right, <strong>and</strong> bottom left panelsrepresent the effect of the ISI of the <strong>in</strong>com<strong>in</strong>g proton, the FSI of the scattered proton, <strong>and</strong> the FSI of thestruck nucleon, respectively; whereas the bottom right figure displays the complete IFSI factor.tributed to the low ejectile k<strong>in</strong>etic energy (T k2 ≈ 114 MeV for the specific case of Fig. 3.7, <strong>and</strong>comparable values for knockout from other levels <strong>and</strong> other nuclei). At such low energies, theRMSGA predictions are not realistic because of the underly<strong>in</strong>g approximations, mostly the postulationof l<strong>in</strong>ear trajectories <strong>and</strong> frozen spectator nucleons. So, for the k<strong>in</strong>ematics discussedhere, the ROMEA method is to be preferred over the RMSGA one, as the latter overestimatesthe distortion for the low-energetic ejectile nucleon.3.2 Radial <strong>and</strong> Polar-Angle Contributions to A(p, pN) Cross SectionsIn Section 3.1.4, it was already mentioned that the IFSI factors are <strong>in</strong>sensitive to the s<strong>in</strong>gleparticlelevel from which the nucleon is ejected. The peculiar spatial characteristics of the differents<strong>in</strong>gle-particle orbits have an impact on the observables, though. Indeed, the distortedmomentum-space wave function φ D α 1of Eq. (2.35) is determ<strong>in</strong>ed by the values of the IFSI factor


Chapter 3. IFSI <strong>and</strong> A(p, pN) Differential Cross Sections 41|S IFSI( 40 Ca)|/|S IFSI( 12 C)|0.80.60.40.2150θ (deg)100500r (fm)1 2 3 4 5 6 7 8Figure 3.5 The ratio of the 40 Ca absolute value of the complete IFSI factor to the 12 C one as a function ofr <strong>and</strong> θ. The results are for proton emission from the Fermi level <strong>and</strong> φ = 0 ◦ , <strong>and</strong> were obta<strong>in</strong>ed with<strong>in</strong>the ROMEA model us<strong>in</strong>g the EDAD1 optical potential of Ref. [46]. K<strong>in</strong>ematics as <strong>in</strong> Fig. 3.1.folded with a relativistic bound-state wave function φ α1 (⃗r). As the particles experience lessIFSI close to the nuclear surface, one obta<strong>in</strong>s a stronger reduction of the quasifree cross sectionfor nucleon knockout from a level which has a larger fraction of its density <strong>in</strong> the nuclear<strong>in</strong>terior. This will become apparent <strong>in</strong> Fig. 3.8, but even more so <strong>in</strong> Section 3.3.1.Fig. 3.8 shows a function δ r (r) which represents the contribution of the nuclear region withradial coord<strong>in</strong>ate r to the differential cross section. The procedure for calculat<strong>in</strong>g this function issimilar to the method exposed <strong>in</strong> Ref. [118] <strong>and</strong> is developed <strong>in</strong> Appendix C. Comparison of theupper <strong>and</strong> lower panels illustrates that IFSI mechanisms make the A(p, 2p) cross sections reflectsurface mechanisms, unlike the A(e, e ′ p) reaction where the weakly <strong>in</strong>teract<strong>in</strong>g electron probesthe entire nuclear volume <strong>and</strong> only the outgo<strong>in</strong>g proton <strong>in</strong>teracts with the residual system.Apart from the shift to higher r, the IFSI br<strong>in</strong>gs about a strong reduction <strong>in</strong> the magnitude ofthe cross sections, whereby the Fermi level is least affected. Even though δ r (r) is concentrated<strong>in</strong> the surface region, the average density seen through this reaction still amounts to 0.069 fm −3(0.080 fm −3 ) or 32% (45%) of the central density <strong>in</strong> the case of 1s 1/2 knockout from 12 C ( 40 Ca).In the case of emission from the Fermi level, on the other h<strong>and</strong>, the average density is only 12%(13%) of the central density for a 12 C ( 40 Ca) target.In Fig. 3.9 we show which (r, θ) coord<strong>in</strong>ates of the collision po<strong>in</strong>t provide the largest contributionsto the cross section. The k<strong>in</strong>ematics was the same as <strong>in</strong> Fig. 3.8 <strong>and</strong> the ROMEAcalculations used the EDAD1 optical potential. As can be <strong>in</strong>ferred from the left panels, thecontributions to the RPWIA cross sections are symmetric around θ = 90 ◦ . The IFSI not onlyshift the maximum <strong>in</strong> δ r,θ to higher r, but also to lower values of θ. This can be expla<strong>in</strong>ed by


3.3. A(p, pN) Differential Cross Sections 42real part of S IFSI10.90.80.70.60.50.40.30.20.10θ = 45 o , φ = 0 o1 2 3 4 5 6r (fm)real part of S IFSI10.90.80.70.60.50.40.30.20.1r = 3 fm, θ = 90 oreal part of S IFSI10.90.80.70.60.50.40.30.20.1φ (deg)r = 4 fm, φ = 0 o00 20 40 60 80 100 120 140 160 180EDAD1EDAD2EDAI00 50 100 150 200 250 300 350θ (deg)Figure 3.6 The sensitivity of the real part of the complete IFSI factor for neutron knockout from 12 Cto the adopted choice for the parametrization of the optical potentials. Results of ROMEA calculationswith the EDAD1 (solid curve), EDAD2 (dashed curve), <strong>and</strong> EDAI (dot-dashed curve) optical potentialsare shown. K<strong>in</strong>ematics as <strong>in</strong> Fig. 3.1.look<strong>in</strong>g at the complete IFSI factor <strong>in</strong> Fig. 3.1: the attenuation effect is clearly larger for θ ≥ 90 ◦ .Apart from break<strong>in</strong>g the symmetry around θ = 90 ◦ , the IFSI also narrow the peak <strong>in</strong> the θdependence. Of course, these IFSI effects are most pronounced for the lowest ly<strong>in</strong>g 1s 1/2 level.3.3 Numerical Results for A(p, pN) Differential Cross Sections3.3.1 12 C, 16 O, <strong>and</strong> 40 Ca(p, 2p) <strong>and</strong> (p, pn) at 1 GeVThe PNPI experiment [114] was carried out with an <strong>in</strong>cident proton beam of energy 1 GeV. Thescattered proton was detected at θ 1 = 13.4 ◦ with a k<strong>in</strong>etic energy between 800 <strong>and</strong> 950 MeV,while the knocked-out nucleon was observed at θ 2 = 67 ◦ hav<strong>in</strong>g a k<strong>in</strong>etic energy below 200 MeV.Figs. 3.10–3.12 display a selection of differential cross section results as a function of thek<strong>in</strong>etic energy of the most energetic nucleon <strong>in</strong> the f<strong>in</strong>al state. The EDAI optical potential [46]was used for the ROMEA calculations. The other parametrizations of Ref. [46] produce similarpredictions, whereas the RMSGA approach fails to give an adequate description of the databecause of the low k<strong>in</strong>etic energy of the ejected nucleon. S<strong>in</strong>ce the experiment of Ref. [114]only measured relative cross sections, the ROMEA results were normalized to the experimentaldata.The ROMEA calculations reproduce the shapes of the measured differential cross sections.


Chapter 3. IFSI <strong>and</strong> A(p, pN) Differential Cross Sections 43real part of S IFSI0.80.60.40.2real part of S IFSI0.80.60.40.2150θ (deg)10050012345r (fm)6150θ (deg)10050012345r (fm)6real part of S IFSI0.80.60.40.20real part of S IFSI0.80.60.40.20150θ (deg)10050012345r (fm)6150θ (deg)10050012345r (fm)6Figure 3.7 As <strong>in</strong> Fig. 3.1, but us<strong>in</strong>g the RMSGA method.Furthermore, comparison of the RPWIA <strong>and</strong> ROMEA calculations shows that the effect of theIFSI is twofold. First, IFSI result <strong>in</strong> a reduction of the RPWIA cross section that is both level <strong>and</strong>A dependent. From the figures it is clear that ejection of a nucleon from a deeper ly<strong>in</strong>g levelleads to stronger <strong>in</strong>itial- <strong>and</strong> f<strong>in</strong>al-state distortions. This reflects the fact that the <strong>in</strong>com<strong>in</strong>g <strong>and</strong>outgo<strong>in</strong>g nucleons encounter more obstacles when a deeper ly<strong>in</strong>g bound nucleon is probed.The A dependence also conforms with our expectations, i.e., the IFSI effects are larger for heaviernuclei. Besides the attenuation, the IFSI also make the measured miss<strong>in</strong>g momentum differentfrom the <strong>in</strong>itial momentum of the struck nucleon. As can be <strong>in</strong>ferred from Fig. 3.13, thismomentum shift leads to an asymmetry between the positive <strong>and</strong> negative miss<strong>in</strong>g-momentumside of the momentum distribution. Note that a positive miss<strong>in</strong>g momentum corresponds top mx = k 1 s<strong>in</strong> θ 1 + k 2 s<strong>in</strong> θ 2 cos φ 2 > 0.3.3.2 40 Ca(p, 2p) at 460 MeVNext, we focus on a 40 Ca(p, 2p) experiment conducted at the EFI facility <strong>in</strong> Chicago by Tyrén etal. [115]. The imp<strong>in</strong>g<strong>in</strong>g proton had a k<strong>in</strong>etic energy of T p1 = 460 MeV, while the outgo<strong>in</strong>g pro-


3.3. A(p, pN) Differential Cross Sections 440.2250.20.17512 C, RPWIA ρ(r)1p 3/21s 1/20.2250.20.17540 Ca, RPWIA ρ(r)1d 3/21s 1/2density (fm -3 )0.150.1250.10.075density (fm -3 )0.150.1250.10.0750.050.02500 1 2 3 4 5 6r (fm)0.050.02500 1 2 3 4 5 6 7 8r (fm)0.2250.20.17512 C, ROMEA EDAD1 ρ(r)1p 3/2 x101s 1/2}0.2250.20.17540 Ca, ROMEA EDAD1 ρ(r)1d 3/2 x151s 1/2}density (fm -3 )0.150.1250.10.075density (fm -3 )0.150.1250.10.0750.050.02500 1 2 3 4 5 6r (fm)0.050.02500 1 2 3 4 5 6 7 8r (fm)Figure 3.8 Contribution to the A(p, 2p) cross section δ r (r) as a function of r. The upper figures presentthe results obta<strong>in</strong>ed after sett<strong>in</strong>g the IFSI factor S IFSI (⃗r) equal to 1 <strong>in</strong> Eq. (C.4), whereas <strong>in</strong> the lowerpanels the ROMEA calculations us<strong>in</strong>g the EDAD1 optical potential of Ref. [46] are depicted. The dashed(dot-dashed) curves show the result for emission from the Fermi (lowest ly<strong>in</strong>g 1s 1/2 ) level. The baryondensity ρ (r) is also shown (solid curve). The ord<strong>in</strong>ate is given for ρ (r). The δ r (r) are plotted <strong>in</strong> units offm 2 up to an arbitrary scal<strong>in</strong>g factor. The k<strong>in</strong>ematics was T p1 = 1 GeV, T k1 = 870 MeV, θ 1 = 13.4 ◦ , <strong>and</strong>θ 2 = 67 ◦ .tons were detected at symmetric coplanar angles (θ 1 = θ 2 <strong>and</strong> φ 2 = 180 ◦ ) <strong>and</strong> equal momenta(k 1 = k 2 ). The differential cross sections for 1d 3/2 , 1d 5/2 , <strong>and</strong> 2s 1/2 knockout are displayed asa function of the scatter<strong>in</strong>g angle θ 1 = θ 2 <strong>in</strong> Figs. 3.14, 3.15, <strong>and</strong> 3.16, respectively. The solidcurves show the ROMEA calculations assum<strong>in</strong>g full occupancy of the shells, while the dashedcurves are scaled by the spectroscopic factors.The shape of the 1d 3/2 cross section is nicely reproduced by the calculations, although thesolid curve somewhat overestimates the magnitude. Normalization to the data leads to a spectroscopicfactor of 0.62. For the 1d 5/2 shell, the overall magnitude of the data is quite wellreproduced, but the shape of the calculated distribution is <strong>in</strong> poor agreement with the data.However, it should be noted that the error bars on the experimental 1d 5/2 cross sections arerelatively large, mak<strong>in</strong>g it difficult to draw any mean<strong>in</strong>gful conclusion. F<strong>in</strong>ally, our calcula-


Chapter 3. IFSI <strong>and</strong> A(p, pN) Differential Cross Sections 45θ (deg)1506θ (deg)15010.810041000.60.4502500.201 2 3 4 5 6r (fm)δ r,θ: 1s 1/2knockout, RPWIA01 2 3 4 5 6r (fm)δ r,θ: 1s 1/2knockout, ROMEAθ (deg)15054θ (deg)15021.5100310021501500.501 2 3 4 5 6r (fm)δ r,θ: 1p 3/2knockout, RPWIA001 2 3 4 5 6r (fm)δ r,θ: 1p 3/2knockout, ROMEA0Figure 3.9 Contribution to the A(p, 2p) cross section δ r,θ (r, θ) (<strong>in</strong> units of fm 2 up to an arbitrary scal<strong>in</strong>gfactor) as a function of the radial coord<strong>in</strong>ate r <strong>and</strong> polar angle θ of the hard collision po<strong>in</strong>t. The functionδ r,θ (r, θ) is def<strong>in</strong>ed <strong>in</strong> Appendix C. The upper (lower) figures show the results for emission from the1s 1/2 (1p 3/2 ) level <strong>in</strong> 12 C.tions describe the shape of the 2s 1/2 cross section very well, but overshoot the data by a factorof more than 2. Therefore, the conclusion that the occupation number for the 2s 1/2 shell issignificantly smaller than the shell-model value of 2 seems <strong>in</strong>escapable. This depletion of theoccupancy of s<strong>in</strong>gle-particle levels that are fully occupied <strong>in</strong> the naive <strong>in</strong>dependent-particlemodel is a general feature of nuclear matter <strong>and</strong> has been confirmed by both experimental <strong>and</strong>theoretical studies [120]. However, the fact that the 2s 1/2 shell is so strongly depleted has so farnot been expla<strong>in</strong>ed theoretically or corroborated by other experiments. For example, from the40 Ca(e, e ′ p) data of Ref. [121], a spectroscopic factor of 0.66 was deduced for the 2s 1/2 orbital.The 1d 3/2 spectroscopic factor of 0.65 obta<strong>in</strong>ed from this experiment, on the other h<strong>and</strong>, doesagree with our value.


3.3. A(p, pN) Differential Cross Sections 46d 5 σ/dΩ 1dΩ 2dT k1(mb/sr 2 MeV)0.90.80.70.60.50.40.30.20.11p 3/2T k1(MeV)ROMEARPWIA/30800 820 840 860 880 900 920 940 960d 5 σ/dΩ 1dΩ 2dT k1(mb/sr 2 MeV)0.60.50.40.30.20.11s 1/2T k1(MeV)ROMEARPWIA/80800 820 840 860 880 900 920 940 960Figure 3.10 Differential cross section for the 12 C(p, 2p) reaction. The solid curve represents the ROMEAcalculation, whereas the dashed curve is the plane-wave result reduced by the <strong>in</strong>dicated factor. TheROMEA results are normalized to the data. Data po<strong>in</strong>ts are from Ref. [114]. The magnitude of theexperimental error bars is estimated to be of the order of 5–10%.d 5 σ/dΩ 1dΩ 2dT k1(mb/sr 2 MeV)0.250.2250.20.1750.150.1250.10.0750.050.0251p 1/2T k1(MeV)ROMEARPWIA/30800 820 840 860 880 900 920 940 960d 5 σ/dΩ 1dΩ 2dT k1(mb/sr 2 MeV)0.350.30.250.20.150.10.051s 1/2T k1(MeV)ROMEARPWIA/80800 820 840 860 880 900 920 940 960Figure 3.11 As <strong>in</strong> Fig. 3.10, but for the 16 O(p, pn) reaction.3.3.3 16 O(p, 2p) <strong>and</strong> (p, pn) at 505 MeVAt TRIUMF, cross sections for scatter<strong>in</strong>g off 16 O <strong>in</strong> coplanar k<strong>in</strong>ematics have been measured atseveral <strong>in</strong>cident energies. First, we consider the 16 O(p, pN) experiment carried out by Mc-Donald et al. [116].Their data were obta<strong>in</strong>ed at an <strong>in</strong>cident energy of 505 MeV with thescattered proton detected at θ 1 = 22.15 ◦ <strong>and</strong> the ejected nucleon at various angles centeredabout the quasifree angle. In Fig. 3.17 our ROMEA calculations are compared with the data for(θ 1 , θ 2 ) = (22.15 ◦ , 53.7 ◦ ).Overall, the shapes of the calculated cross sections as a function of the k<strong>in</strong>etic energy T k1 ofthe scattered proton follow the trend of the experimental results. In general, the experimentalcross sections for knockout from the 1p 1/2 state are described somewhat better than the 1p 3/2data. Further, the calculations overestimate the experimental values by a factor between 1.2 <strong>and</strong>1.8. The spectroscopic factors normaliz<strong>in</strong>g the calculations to the data are the largest (smallest)for neutron knockout from the 1p 3/2 shell (proton knockout from the 1p 1/2 shell) <strong>and</strong> depend


Chapter 3. IFSI <strong>and</strong> A(p, pN) Differential Cross Sections 47d 5 σ/dΩ 1dΩ 2dT k1(mb/sr 2 MeV)1.41.210.80.60.40.22s 1/2T k1(MeV)ROMEARPWIA/60800 820 840 860 880 900 920 940 960d 5 σ/dΩ 1dΩ 2dT k1(mb/sr 2 MeV)0.250.2250.20.1750.150.1250.10.0750.050.0251d 5/2T k1(MeV)ROMEARPWIA/100800 820 840 860 880 900 920 940 960Figure 3.12 As <strong>in</strong> Fig. 3.10, but for the 40 Ca(p, pn) reaction.1201001p 3/2ROMEARPWIAρ(p m) (GeV -3 )806040200-300 -200 -100 0 100 200 300p m(MeV/c)Figure 3.13 The 12 C(p, 2p) momentum distribution for the 1p 3/2 state as a function of the miss<strong>in</strong>g momentum.The solid (dashed) curve represent ROMEA (RPWIA) calculations.slightly on the choice for the optical potential. The results of our calculations are similar tothose of Ref. [31].3.3.4 16 O(p, 2p) at 200 MeVAnother TRIUMF experiment [8] studied the reaction 16 O(p, 2p) at 200 MeV <strong>in</strong>com<strong>in</strong>g energy.The differential cross section d 5 σ/dΩ 1 dΩ 2 dE diff , where E diff = T k1 − T k2 , was measured at10 pairs of coplanar scatter<strong>in</strong>g angles (θ 1 , θ 2 ). Fig. 3.18 shows a representative sampl<strong>in</strong>g ofthe cross-section results. Overall, the ROMEA model provides a good description of the dataas both the peak positions <strong>and</strong> the widths of the experimental cross sections are reproduced.Aga<strong>in</strong>, the agreement with the data is slightly better for the 1p 1/2 shell. The values for thespectroscopic factors are consistent with those of Section 3.3.3.


3.3. A(p, pN) Differential Cross Sections 48d 5 σ/dΩ 1dΩ 2dT k1(µb/sr 2 MeV)1601401d 3/212010080604020020 25 30 35 40 45 50 55 60θ 1= θ 2(deg)Figure 3.14 Symmetric coplanar angular distribution for the 40 Ca(p, 2p) 39 K(1d −13/2) reaction at 460 MeV.The curves are obta<strong>in</strong>ed from a ROMEA calculation with the EDAD2 optical potential [46]. The data aretaken from Ref. [119], a re-analysis of the orig<strong>in</strong>al experiment [115].d 5 σ/dΩ 1dΩ 2dT k1(µb/sr 2 MeV)1601401d 5/212010080604020020 25 30 35 40 45 50 55 60θ 1= θ 2(deg)Figure 3.15 As <strong>in</strong> Fig. 3.14, but for knockout from the 1d 5/2 shell.3.3.5 4 He(p, 2p) at 250 MeVF<strong>in</strong>ally, we present some results for the 4 He(p, 2p) reaction at an <strong>in</strong>cident proton energy of250 MeV. Fig. 3.19 compares the data from the TRIUMF experiment of Ref. [9] with ROMEAcalculations us<strong>in</strong>g the optical potential of van Oers et al. [9]. The typical shape for knockout ofan s-state proton is reproduced by the ROMEA predictions.The analysis of the different experiments <strong>in</strong> this chapter has shown that our ROMEA formalismexpla<strong>in</strong>s the global shape of the differential cross sections. The fair agreement with thedata of Sections 3.3.4 <strong>and</strong> 3.3.5 demonstrates that the ROMEA model works satisfactorily atlower <strong>in</strong>cident energies. While the shape of the A(p, pN) differential cross sections is probablynot substantially affected by sp<strong>in</strong> effects, this is not the case for A(p, pN) sp<strong>in</strong> observableslike the analyz<strong>in</strong>g power. To adequately describe these observables, it is necessary to add the


Chapter 3. IFSI <strong>and</strong> A(p, pN) Differential Cross Sections 49d 5 σ/dΩ 1dΩ 2dT k1(µb/sr 2 MeV)160014002s 1/212001000800600400200020 25 30 35 40 45 50 55 60θ 1= θ 2(deg)Figure 3.16 As <strong>in</strong> Fig. 3.14, but for knockout from the 2s 1/2 shell.sp<strong>in</strong>-orbit potential <strong>in</strong> the treatment of the IFSI <strong>in</strong> our model.


3.3. A(p, pN) Differential Cross Sections 50d 5 σ/dΩ 1dΩ 2dT k1(mb/sr 2 MeV)0.20.180.160.140.120.10.080.060.040.0216 O(p,2p) 15 N1p 1/2d 5 σ/dΩ 1dΩ 2dT k1(mb/sr 2 MeV)0.20.180.160.140.120.10.080.060.040.0216 O(p,2p) 15 N1p 3/20300 320 340 360 380 400 420 440T k1(MeV)0300 320 340 360 380 400 420 440T k1(MeV)d 5 σ/dΩ 1dΩ 2dT k1(mb/sr 2 MeV)0.120.10.080.060.040.0216 O(p,pn) 15 O1p 1/2d 5 σ/dΩ 1dΩ 2dT k1(mb/sr 2 MeV)0.20.180.160.140.120.10.080.060.040.0216 O(p,pn) 15 O1p 3/20300 320 340 360 380 400 420 440T k1(MeV)0300 320 340 360 380 400 420 440T k1(MeV)Figure 3.17 Differential cross sections for the 16 O(p, 2p) 15 N (upper row) <strong>and</strong> 16 O(p, pn) 15 O (lower row)reactions. The left panels correspond with knockout from the 1p 1/2 orbital, while the right panels show1p 3/2 -shell knockout cross sections. The <strong>in</strong>com<strong>in</strong>g proton beam had an energy of 505 MeV <strong>and</strong> thescatter<strong>in</strong>g angles of the outgo<strong>in</strong>g nucleons were (θ 1 , θ 2 ) = (22.15 ◦ , 53.7 ◦ ). The curves are ROMEA calculationswith the EDAD1 optical potential [46], normalized to the data. The data are from Ref. [116].


Chapter 3. IFSI <strong>and</strong> A(p, pN) Differential Cross Sections 5180θ 1= 30 o , θ 2= 45 o80θ 1= 30 o , θ 2= 50 od 5 σ/dΩ 1dΩ 2dE diff(µb/sr 2 MeV)706050403020101p 3/21p 1/2d 5 σ/dΩ 1dΩ 2dE diff(µb/sr 2 MeV)706050403020101p 3/21p 1/2040 60 80 100 120 140T k1(MeV)040 60 80 100 120 140T k1(MeV)80θ 1= 30 o , θ 2= 65 o6θ 1= 65 o , θ 2= 65 od 5 σ/dΩ 1dΩ 2dE diff(µb/sr 2 MeV)706050403020101p 3/21p 1/2d 5 σ/dΩ 1dΩ 2dE diff(µb/sr 2 MeV)543211p 3/21p 1/2040 60 80 100 120 140T k1(MeV)040 60 80 100 120 140T k1(MeV)Figure 3.18 16 O(p, 2p) differential cross sections for four of the angle comb<strong>in</strong>ations (θ 1 , θ 2 ) of experiment[8]. The curves show<strong>in</strong>g our ROMEA calculations based on the EDAI optical potential [46] arenormalized to the data po<strong>in</strong>ts from Ref. [8]. The solid l<strong>in</strong>es <strong>and</strong> red circles perta<strong>in</strong> to the 1p 3/2 state,while the dashed l<strong>in</strong>es <strong>and</strong> green triangles correspond to the 1p 1/2 state.


3.3. A(p, pN) Differential Cross Sections 52d 5 σ/dΩ 1dΩ 2dT k1(mb/sr 2 MeV)1.210.80.60.40.2θ 1= 40 oθ 2= 40 oROMEARPWIAd 5 σ/dΩ 1dΩ 2dT k1(mb/sr 2 MeV)1.210.80.60.40.2θ 1= 45 oθ 2= 35 o T k1(MeV)070 80 90 100 110 120 130 140 150 160T k1(MeV)030 40 50 60 70 80 90 100 110 120Figure 3.19 Differential cross section for the 4 He(p, 2p) reaction at angle pairs (40 ◦ , 40 ◦ ) <strong>and</strong> (45 ◦ , 35 ◦ )at 250 MeV. The solid (dashed) curves refer to ROMEA (RPWIA) results. The data are from Ref. [9].


Chapter 4<strong>Nuclear</strong> Transparency from A(p, 2p)ReactionsThe transition region between nucleon-meson (hadronic) <strong>and</strong> quark-gluon (partonic) degreesof freedom is a topic of longst<strong>and</strong><strong>in</strong>g <strong>in</strong>terest <strong>in</strong> nuclear physics. A promis<strong>in</strong>g observable tomap this transition is the transparency of the nuclear medium to the propagation of hadrons.In A(p, 2p) experiments, the nuclear transparency is def<strong>in</strong>ed as the ratio of the cross sectionper nucleon to the hydrogen one. Accord<strong>in</strong>gly, the nuclear transparency is a measure for theattenuation effects of the spectator nucleons on the imp<strong>in</strong>g<strong>in</strong>g <strong>and</strong> outgo<strong>in</strong>g protons.In the conventional Glauber picture [71], the nuclear transparency extracted from A(p, 2p)reactions is predicted to be rather constant for <strong>in</strong>com<strong>in</strong>g momentum larger than a few GeV/c.The color transparency (CT) phenomenon suggests an anomalously large transmission probabilityof protons through nuclei [122, 123] <strong>and</strong> leads to a nuclear transparency that <strong>in</strong>creaseswith <strong>in</strong>com<strong>in</strong>g momentum. At Brookhaven National Laboratory (BNL), two experiments,E834 [124] <strong>and</strong> E850 [125, 126], measured the nuclear transparency <strong>in</strong> the A(p, 2p) quasielasticscatter<strong>in</strong>g process near 90 ◦ <strong>in</strong> the pp c.m. frame. As can be appreciated from Fig. 4.1, thedata suggest a CT-like <strong>in</strong>crease <strong>in</strong> the transparency for imp<strong>in</strong>g<strong>in</strong>g proton momenta between 5<strong>and</strong> 10 GeV/c. For higher momenta, the measured nuclear transparency falls back to the levelpredicted by typical Glauber calculations. This oscillatory energy dependence is not unique tothe A(p, 2p) nuclear transparency: it has been observed or h<strong>in</strong>ted at <strong>in</strong> pp elastic scatter<strong>in</strong>g [127],elastic πp fixed-angle scatter<strong>in</strong>g [128–134], pion photoproduction [134–137], <strong>and</strong> deuteron photodis<strong>in</strong>tegration[138–143].Two explanations for the measured energy dependence of the A(p, 2p) transparency havebeen proposed [144,145]. Both the Ralston-Pire [144] <strong>and</strong> the Brodsky-de Teramond model [145]are based on the presence of two terms <strong>in</strong> the free pp scatter<strong>in</strong>g amplitude: one represent<strong>in</strong>g asmall object <strong>and</strong> the other represent<strong>in</strong>g a normal-sized object. The <strong>in</strong>terference between thesetwo amplitudes <strong>in</strong>duces the oscillation of the free pp cross section about the scal<strong>in</strong>g behavior.Inside the nuclear medium, the normal-sized components <strong>in</strong> the hadron wave functions arefiltered away. This phenomenon of “nuclear filter<strong>in</strong>g” (NF) [144, 146, 147] reproduces qualita-53


54Figure 4.1 The beam-momentum dependence of the A(p, 2p) nuclear transparency as measured <strong>in</strong> theE834 experiment [124]. The upper panel shows the transparency as a function of the <strong>in</strong>com<strong>in</strong>g lab momentump 1 for various nuclear targets <strong>and</strong> −0.2 GeV/c < p mz < 0.1 GeV/c, with p mz the longitud<strong>in</strong>alcomponent of the miss<strong>in</strong>g momentum. In the bottom panel, the 27 Al data po<strong>in</strong>ts are plotted versus theeffective <strong>in</strong>cident momentum p eff = p 1 (1 − p mz /M p ), which takes the motion of the target proton <strong>in</strong> thenucleus <strong>in</strong>to account. The dashed l<strong>in</strong>e <strong>in</strong> the bottom panel corresponds to a Glauber calculation. Thisfigure is taken from Ref. [124].tively the observed bump <strong>in</strong> the A(p, 2p) transparency.In this chapter, the relativistic <strong>and</strong> cross-section factorized framework of Section 2.1 is extendedto <strong>in</strong>corporate the Ralston-Pire model for the pp scatter<strong>in</strong>g amplitude <strong>and</strong> the conceptsof CT <strong>and</strong> NF. Section 4.1 is devoted to the Ralston-Pire description of elastic pp scatter<strong>in</strong>g. InSection 4.2, the idea of nuclear filter<strong>in</strong>g is reviewed <strong>and</strong> the A(p, 2p) formalism of Section 2.1is adjusted to take <strong>in</strong>to account both the small-sized <strong>and</strong> the normal-sized contribution to thehard pp scatter<strong>in</strong>g amplitude. The IFSI effects are computed with<strong>in</strong> the RMSGA approaches ofSections 2.3 <strong>and</strong> 2.4. The phenomenon of CT constitutes the subject of Section 4.3. To estimateits effects, we consider the quantum diffusion model of Ref. [148] <strong>and</strong> the hadronic picture ofJenn<strong>in</strong>gs <strong>and</strong> Miller [149]. The comparison between the two CT models is made <strong>in</strong> a consistentway. This implies that all the <strong>in</strong>gredients <strong>in</strong> the transparency calculations not related toCT are kept identical. Section 4.4 presents the nuclear transparency results for the target nuclei7 Li, 12 C, 27 Al, <strong>and</strong> 63 Cu. The effect of IFSI, CT, <strong>and</strong> NF is discussed. Furthermore, the


Chapter 4. <strong>Nuclear</strong> Transparency from A(p, 2p) Reactions 55accuracy of the approximated RMSGA approach is <strong>in</strong>vestigated <strong>and</strong> the two CT treatments arecompared. F<strong>in</strong>ally, some conclud<strong>in</strong>g remarks <strong>and</strong> an outlook with regard to the research <strong>in</strong>tonuclear transparency can be found <strong>in</strong> Section 4.5.The results conta<strong>in</strong>ed <strong>in</strong> this chapter have been published <strong>in</strong> Ref. [150].4.1 High-Momentum-Transfer Wide-Angle Scatter<strong>in</strong>gAt low energies or long distances, the global features of the strong <strong>in</strong>teraction can be describedwith<strong>in</strong> the nucleon-meson picture [151]. At high energies or short distances, perturbative QCD(pQCD) provides a precise description of hadronic reactions <strong>in</strong> terms of quark-gluon degreesof freedom. At present, it is not clear how these two regimes are connected. Exclusive hadronicreactions such as proton-proton elastic scatter<strong>in</strong>g, meson photoproduction <strong>and</strong> deuteron photodis<strong>in</strong>tegrationhave been studied extensively <strong>in</strong> order to explore this transition region. Excellentreviews can be found <strong>in</strong> Refs. [152, 153].4.1.1 Quark-Count<strong>in</strong>g RuleAt high energy <strong>and</strong> large momentum transfer, the differential cross section for many exclusiveA + B → C + D reactions [154] obeys the quark-count<strong>in</strong>g (QC) rule [155–158]( ) dσ∝ s 2−n f( t dts ) , (4.1)AB→CDtfor s → ∞ ,s ≡ const , n = n A + n B + n C + n D .Here, the variables s <strong>and</strong> t are the M<strong>and</strong>elstam variabless = (P A + P B ) 2 , t = (P A − K C ) 2 , (4.2)where P A , P B <strong>and</strong> P C are the four-momenta of the hadrons. n A , n B , n C , <strong>and</strong> n D are the numberof valence quarks <strong>in</strong>side the hadrons A, B, C, <strong>and</strong> D, respectively. For pp scatter<strong>in</strong>g, thedifferential cross section scales roughly as ( )dσdt pp→pp ∝ s−10 . As for the function f( t s), whichdescribes the angular dependence of the differential cross section, no rigorous pQCD calculationhas been performed yet. The constituent <strong>in</strong>terchange model [159] predictsf( t s ) ≡ f(cos θ c.m.) = ( 1 − cos 2 θ c.m.) −4γ, (4.3)where γ can be <strong>in</strong> the range 1.3–2.0.Eq. (4.1) is also known as the dimensional scal<strong>in</strong>g law, s<strong>in</strong>ce it was orig<strong>in</strong>ally derived [155–157] us<strong>in</strong>g dimensional analysis. The ma<strong>in</strong> assumption of this analysis is that only one constituent(a dimensionless po<strong>in</strong>t-like quark) of each hadron participates <strong>in</strong> the reaction whereasthe other constituents are spectators. In order for the spectator quarks to be able to follow the<strong>in</strong>teract<strong>in</strong>g quarks after the wide-angle scatter<strong>in</strong>g, each hadron must fluctuate to its m<strong>in</strong>imalFock space component dur<strong>in</strong>g the reaction. This m<strong>in</strong>imal state must be much smaller than a


4.1. High-Momentum-Transfer Wide-Angle Scatter<strong>in</strong>g 56regular hadron because otherwise the spectator quarks will cont<strong>in</strong>ue <strong>in</strong> their orig<strong>in</strong>al direction<strong>in</strong>dependently from the <strong>in</strong>teract<strong>in</strong>g quark <strong>and</strong> produce more hadrons, thus destroy<strong>in</strong>g the exclusivenature of the reaction. The amplitude for such a fluctuation is per) ni −1hadron,where q is the momentum transfer of the reaction (assumed to be transferred directly betweenthe two <strong>in</strong>teract<strong>in</strong>g quarks) <strong>and</strong> n i is the number of quarks <strong>in</strong> hadron i. The comb<strong>in</strong>ation ofthese amplitudes leads to Eq. (4.1). Later, this naive description was confirmed by a shortdistancepQCD derivation [158].Fig. 4.2 shows the diagrams which contribute to the quark-count<strong>in</strong>g scatter<strong>in</strong>g. The QCprocess is characterized by the exchange of five hard gluons <strong>in</strong>side the <strong>in</strong>teract<strong>in</strong>g hadrons.The quark-<strong>in</strong>terchange diagram presented <strong>in</strong> the right panel of Fig. 4.2 dom<strong>in</strong>ates the crosssection [160, 161].(1qFigure 4.2 Typical diagrams for quark-count<strong>in</strong>g scatter<strong>in</strong>g. The left panel shows the pure gluonexchangediagram, while the quark-<strong>in</strong>terchange mechanism is represented <strong>in</strong> the right panel.Despite the success of pQCD <strong>in</strong> describ<strong>in</strong>g the energy dependence of exclusive cross sectionsat fixed center-of-mass scatter<strong>in</strong>g angle, there are strong <strong>in</strong>dications that the short-distancepQCD picture might be wrongly built <strong>in</strong>to dogma. The hadron helicity conservation (HHC)rules predicted by the asymptotic short-distance approach of Brodsky <strong>and</strong> Lepage [162] areviolated by experimental data <strong>in</strong> the same energy <strong>and</strong> momentum region [163–166]. However,the statement that hadron helicity flip is <strong>in</strong> contradiction with pQCD is currently underdebate [167, 168]. In pp scatter<strong>in</strong>g, the strik<strong>in</strong>g energy <strong>and</strong> angular dependence of the sp<strong>in</strong>sp<strong>in</strong>correlation factor A NN attracts attention [169–173]. Furthermore, pQCD calculations ofthe magnitudes of form factors [174–177] fail to describe the data, as non-perturbative longdistanceeffects are shown to play a crucial role. Often, the QC rule of Eq. (4.1) is only approximatelyobeyed. Indeed, when exam<strong>in</strong>ed <strong>in</strong> detail, several reactions exhibit oscillationsaround the overall scal<strong>in</strong>g-law dependence. The oscillatory behavior of the free pp scatter<strong>in</strong>g isillustrated <strong>in</strong> Fig. 4.3.


Chapter 4. <strong>Nuclear</strong> Transparency from A(p, 2p) Reactions 57Figure 4.3 The energy dependence of R 1 (s) = const s 10 (dσ/dt) pp→pp| 90 ◦ for the high-energy pp elasticscatter<strong>in</strong>g at 90 ◦ c.m. angle. The data are compared to the parametrization of Eq. (4.8) (solid l<strong>in</strong>e). Thisfigure is taken from Ref. [178].4.1.2 L<strong>and</strong>shoff Scatter<strong>in</strong>gBesides the quark-count<strong>in</strong>g scatter<strong>in</strong>g of Section 4.1.1, another scatter<strong>in</strong>g mechanism is consideredto play a role <strong>in</strong> exclusive high-momentum-transfer hadronic reactions. This process wasproposed by L<strong>and</strong>shoff [179–181] <strong>and</strong> is referred to as the multiple-scatter<strong>in</strong>g model, <strong>in</strong>dependentscatter<strong>in</strong>g, or L<strong>and</strong>shoff scatter<strong>in</strong>g. The diagram for this process is displayed <strong>in</strong> Fig. 4.4:each constituent quark of one hadron scatters on one of the quarks of the other hadron. Thepairs of quarks scatter <strong>in</strong>dependently <strong>and</strong> the energy dependence of the cross section is obta<strong>in</strong>edthrough the condition that the outgo<strong>in</strong>g quarks must align to recomb<strong>in</strong>e to a hadron.The s dependence of the cross section is found to be( ) dσ∝ s −n f( t dts ) , (4.4)AB→CDtfor s → ∞ , ≪ 1 , n = 8 for pp scatter<strong>in</strong>g .sThe ma<strong>in</strong> feature of this picture is that the hadrons stay normal-sized dur<strong>in</strong>g the <strong>in</strong>teraction.At first sight, the L<strong>and</strong>shoff mechanism is expected to dom<strong>in</strong>ate at high energies, <strong>in</strong> disagreementwith the available data. However, L<strong>and</strong>shoff scatter<strong>in</strong>g is partially suppressed by softgluon radiation [182]. The radiative corrections are taken <strong>in</strong>to account through Sudakov formfactors [183–186] <strong>and</strong> br<strong>in</strong>g the cross section dependence for the L<strong>and</strong>shoff mechanism close todσ/dt ∝ s −10 for pp scatter<strong>in</strong>g [178].4.1.3 Ralston-Pire Picture <strong>and</strong> Parametrization of the Free pp Cross SectionRalston <strong>and</strong> Pire expla<strong>in</strong>ed the oscillations <strong>in</strong> free pp scatter<strong>in</strong>g by the <strong>in</strong>terference between thequark-count<strong>in</strong>g <strong>and</strong> the L<strong>and</strong>shoff amplitude. In their approach [178,187], the sp<strong>in</strong>-averaged pp


4.1. High-Momentum-Transfer Wide-Angle Scatter<strong>in</strong>g 58Figure 4.4 Scatter<strong>in</strong>g via the L<strong>and</strong>shoff mechanism.scatter<strong>in</strong>g amplitude consists of the quark-count<strong>in</strong>g (QC) <strong>and</strong> the L<strong>and</strong>shoff (L) contribution:M pp = M ppQC + Mpp L . (4.5)The L<strong>and</strong>shoff term can be related to the quark-count<strong>in</strong>g term throughM ppL = ρ √( )1 s2 1 GeV 2 e± i φ(s)+δ 1M ppQC . (4.6)Here, ρ 1 = 0.08 <strong>and</strong> δ 1 = −2.0 is an <strong>in</strong>calculable, energy-<strong>in</strong>dependent phase. These valueswere determ<strong>in</strong>ed from a fit to the pp data at 90 ◦ [178, 187]. The energy-dependent phase φ (s)arises from the gluonic radiative corrections <strong>and</strong> is calculable <strong>in</strong> pQCD [178, 187, 188]:φ (s) =π { [0.06 ln s]}ln0.01 GeV 2 . (4.7)The quantum electrodynamics (QED) variant of this effect is observed <strong>in</strong> charged-particle scatter<strong>in</strong>g:the Coulomb-nuclear <strong>in</strong>terference phase shift between the strong-<strong>in</strong>teraction <strong>and</strong> theelectromagnetic contributions to the scatter<strong>in</strong>g amplitude [189–191].By analogy with thisCoulomb-nuclear <strong>in</strong>terference effect, φ (s) is called the chromo-Coulomb phase shift. The soleparameter which rema<strong>in</strong>s undeterm<strong>in</strong>ed is the sign of the phase difference φ (s) + δ 1 betweenthe quark-count<strong>in</strong>g <strong>and</strong> the L<strong>and</strong>shoff term. Therefore, both possibilities are tested <strong>in</strong> the calculationsof the nuclear transparency.For the parametrization of the free pp scatter<strong>in</strong>g cross section ( )dσdt, the Ralston-Pirepp→ppseparation of Eq. (4.6) is comb<strong>in</strong>ed with the θ c.m. dependence suggested by [159] (see Eq. (4.3)).This leads to [67]( ) dσdtpp→pp∝∣∣M pp= 45.0×∣QC + Mpp L ∣ 2( )µb 10 GeV2 10(1 − cos 2 ) −4γsr GeV 2 θ c.m.s[ √ s1 + ρ 1GeV 2 cos φ (s) + ρ2 14sGeV 2 ]F (s, θ c.m. ) , (4.8)


Chapter 4. <strong>Nuclear</strong> Transparency from A(p, 2p) Reactions 59where γ = 1.6. The function F (s, θ c.m. ) is used to further adjust the phenomenologically motivatedparametrization of the experimental data <strong>in</strong> the range 60 ◦ ≤ θ c.m. ≤ 90 ◦ [192].On a log-log plot, the pp → pp data fall roughly like s −10 , but a closer look reveals thatthe scal<strong>in</strong>g-law behavior is modulated by oscillations with the logarithm of the energy, as wasshown <strong>in</strong> Fig. 4.3.4.1.4 Brodsky-de Teramond PictureAnother <strong>in</strong>terpretation of the behavior of the pp cross section <strong>and</strong> sp<strong>in</strong>-sp<strong>in</strong> correlation wasgiven by Brodsky <strong>and</strong> de Teramond [145]. In their model, two broad baryon resonances areassociated with strange <strong>and</strong> charmed particle production thresholds at √ s = 2.55 <strong>and</strong> 5.08 GeV.The resonances represent normal-sized hadrons <strong>and</strong> <strong>in</strong>terfere with the small-sized quark-count<strong>in</strong>gamplitude, thereby expla<strong>in</strong><strong>in</strong>g the anomalous behavior of the pp cross section, the pp sp<strong>in</strong>sp<strong>in</strong>correlation, <strong>and</strong> the A(p, 2p) transparency. Their idea is supported by the fact that themass scale of the charm threshold concurs with that of the rapid energy variation <strong>in</strong> the sp<strong>in</strong>sp<strong>in</strong>correlation <strong>and</strong> nuclear transparency data. In our calculations, we restrict ourselves to themodel of Ralston <strong>and</strong> Pire.4.2 <strong>Nuclear</strong> Filter<strong>in</strong>gInside the nuclear medium, the normal-sized component (the L<strong>and</strong>shoff amplitude) is suppressedby the strong <strong>in</strong>teractions with the spectator nucleons, while the small-sized component(the quark-count<strong>in</strong>g amplitude) escapes the nucleus with relatively small attenuation dueto CT (see Section 4.3). This phenomenon is called “nuclear filter<strong>in</strong>g” [144, 146, 147]: the nucleusfilters away the normal-sized components. Accord<strong>in</strong>gly, the nucleus plays an active role<strong>in</strong> select<strong>in</strong>g small-sized components. As a result, the oscillations seen <strong>in</strong> free pp scatter<strong>in</strong>g willbe weaker <strong>in</strong> A(p, 2p) reactions.Incorporat<strong>in</strong>g the Ralston-Pire approach <strong>and</strong> the NF mechanism <strong>in</strong>to the A(p, 2p) formalismof Section 2.1, the amplitude for the p (E p1 , ⃗p 1 , m s1i ) + A(E A , ⃗ )k A , 0 + → p(E k1 , ⃗ )k 1 , m s1f +p(E k2 , ⃗ )k 2 , m s2f + A − 1(E A−1 , ⃗ )k A−1 , J R M R reaction becomesM (p,2p)fi= ∑ ( )M ppQCmm s1i ,m s,m s1f ,m s2fsū(⃗p m , m s ) φ RMSGA+CTα 1(⃗p m )+ ∑ (MppL)m s1i ,m s,m ū(⃗ps1f ,m m , m s ) φ RMSGAs2f α 1(⃗p m ) , (4.9)m swhere ⃗p m is the miss<strong>in</strong>g momentum <strong>and</strong> α 1 refers to the state where<strong>in</strong> the struck protonresided. The distorted momentum-space wave functions φα RMSGA(+CT)1(⃗p m ) are def<strong>in</strong>ed as <strong>in</strong>Eq. (2.35). The IFSI of the imp<strong>in</strong>g<strong>in</strong>g <strong>and</strong> two outgo<strong>in</strong>g protons are calculated <strong>in</strong> the RMSGAframeworks of Sections 2.3 <strong>and</strong> 2.4. S<strong>in</strong>ce the quark-count<strong>in</strong>g term is associated with smallsize, the correspond<strong>in</strong>g momentum-space wave function φ RMSGA+CTα 1(⃗p m ) <strong>in</strong>cludes the effect


4.3. Color Transparency 60of CT. The L<strong>and</strong>shoff term, on the other h<strong>and</strong>, corresponds with a hadron of normal size. Consequently,the IFSI can be computed <strong>in</strong> st<strong>and</strong>ard Glauber theory.Us<strong>in</strong>g the sp<strong>in</strong>-averaged pp matrix element of Eq. (4.5), the squared A(p, 2p) matrix elementfor knockout from the α 1 shell can be cast <strong>in</strong> the form∑∣if∣M (p,2p)fi∣ 2 = ∑ { ∣∣∣M pp∣ 2 ∣ ∣ū(⃗p m , m s ) φ RMSGA+CTα 1(⃗p m ) ∣ 2m,m s[+2Re M ppQCQC(Mpp) ∗L ū(⃗pm , m s ) φ RMSGA+CT+ ∣ ∣M ppLα 1(⃗p m )× ( ū(⃗p m , m s ) φ RMSGAα 1(⃗p m ) ) ]∗∣ 2 ∣ ∣ū(⃗p m , m s ) φ RMSGAα 1(⃗p m ) ∣ }2, (4.10)with m the struck nucleon’s generalized angular momentum quantum number. The differentialcross section is obta<strong>in</strong>ed as an <strong>in</strong>coherent sum of the squared matrix elements over all protonlevels α 1 , thereby factor<strong>in</strong>g <strong>in</strong> the occupation number of every level.For all the results presented <strong>in</strong> this chapter, the relativistic bound-state wave function φ α1 (⃗r)is computed <strong>in</strong> the Hartree approximation to the σ − ω model [38], us<strong>in</strong>g the W1 parametrizationfor the different field strengths [193] (see Appendix B). Hereafter, results obta<strong>in</strong>ed on thebasis of Eq. (4.10) are dubbed RMSGA + CT + NF.In our numerical calculations, we will also consider the st<strong>and</strong>ard RMSGA + CT picture.In this scenario, the entire wave packet of the <strong>in</strong>com<strong>in</strong>g <strong>and</strong> outgo<strong>in</strong>g protons is assumed tobe <strong>in</strong> a small-sized configuration, which propagates through a passive nuclear medium. Thisamounts to neglect<strong>in</strong>g the L<strong>and</strong>shoff term <strong>in</strong> the amplitudes M pp <strong>and</strong> M (p,2p)fiof Eqs. (4.5) <strong>and</strong>(4.9). F<strong>in</strong>ally, <strong>in</strong> the st<strong>and</strong>ard RMSGA calculations, both the L<strong>and</strong>shoff term <strong>and</strong> CT effects areneglected.4.3 Color Transparency4.3.1 Requirements for CTThe manifestation of the quark-gluon degrees of freedom gives rise to several phenomena <strong>in</strong>exclusive hadronic reactions. One such prediction of QCD is color transparency [122, 123].The basic idea of CT is that hadrons, which <strong>in</strong>teract very strongly with the nuclear mediumunder normal circumstances, form small objects (so-called po<strong>in</strong>t-like configurations, or PLCs)that move through nuclear matter almost undisturbed. So, while NF uses the nuclear mediumactively, <strong>in</strong> CT large momentum transfer selects short-distance objects <strong>and</strong> the nucleus functionsas a passive medium for the passage of the PLCs. The physics of CT is based on threeideas [194–197]:


Chapter 4. <strong>Nuclear</strong> Transparency from A(p, 2p) Reactions 61(i) Small objects are produced <strong>in</strong> high-momentum-transfer reactionsAs can be seen <strong>in</strong> Fig. 2.4, the nucleon-nucleon <strong>in</strong>teraction becomes predom<strong>in</strong>antly <strong>in</strong>elasticat high energies. Consequently, a nucleon struck by a highly energetic particle (hadron, lepton,. . . ) is expected to scatter <strong>in</strong>to the <strong>in</strong>elastic channels. However, due to the association ofhigh momentum transfer with small wavelengths, the imp<strong>in</strong>g<strong>in</strong>g particle can probe the subnucleonicdegrees of freedom. Suppose now that the <strong>in</strong>com<strong>in</strong>g particle hits one of the conf<strong>in</strong>edcolored quarks. The excited quark becomes off-shell with an energy surplus of δE ∼ ω <strong>and</strong> willdecay with a lifetime τ ∼ 1/ω by emitt<strong>in</strong>g gluons. Ow<strong>in</strong>g to quark conf<strong>in</strong>ement, two th<strong>in</strong>gscan happen: the colored quark hadronizes or it recomb<strong>in</strong>es with its two companion quarks. Inorder for the second option to take place, the quarks must have been close together <strong>in</strong>itially.The radiated gluons must be absorbed by quarks which were at most a distance r ∼ 1/ω awayfrom the off-shell quark. This implies that the struck nucleon must have been <strong>in</strong> a small-sizedfluctuation or po<strong>in</strong>t-like configuration.QCD lattice calculations <strong>in</strong>dicate that hadrons are bound states of strongly <strong>in</strong>teract<strong>in</strong>g quarks<strong>and</strong> gluons. Each hadron can be characterized <strong>in</strong> terms of an <strong>in</strong>f<strong>in</strong>ite number of configurations,e.g., |qqq〉, |qqq + π〉, . . . with vary<strong>in</strong>g sizes. A quantum system fluctuates between its differentconfigurations. Hence, “snapshots” of a hadron taken at different times would disclose bothsmall- <strong>and</strong> large-sized configurations, a phenomenon referred to as color fluctuations. On thebasis of the uncerta<strong>in</strong>ty pr<strong>in</strong>ciple, the time scale for fluctuations can be estimated to be <strong>in</strong>verselyproportional to the mass difference of the two configurations: τ ∼ 1/(m − M). The relevantmass differences are typically of the order of hundreds of MeV, so the fluctuation time is of theorder of 1 fm/c.(ii) Small objects experience reduced <strong>in</strong>teractionsThe second requirement on which the existence of CT depends, is that a small object <strong>in</strong>teracts<strong>in</strong> an anomalously weak manner with the surround<strong>in</strong>g nuclear medium. This phenomenon ofcolor neutrality or color screen<strong>in</strong>g is the QCD analogue of the charge screen<strong>in</strong>g effect <strong>in</strong> QED<strong>and</strong> arises naturally from a two-gluon-exchange model between color s<strong>in</strong>glets [198–200]. Asgluons carry color, s<strong>in</strong>gle gluon exchange is forbidden. A small object has a small color dipolemoment <strong>and</strong> <strong>in</strong>teracts with nuclear matter <strong>in</strong> a much weaker way than a normal-sized object.In this respect, the force between two color s<strong>in</strong>glets can be regarded as a “color Van der Waalsforce”.(iii) Small objects can be considered as frozen while travers<strong>in</strong>g the nuclear mediumA PLC is not a stationary eigenstate of the QCD Hamiltonian, but a superposition of eigenstates.As such, it is subject to time evolution, lead<strong>in</strong>g to an <strong>in</strong>crease of its transverse size<strong>and</strong> the restoration of soft quark-gluon fields [148, 201]. Suppose a hadron with a large laboratorymomentum p fluctuates from a PLC with bare mass M to a normal-sized hadron


4.3. Color Transparency 62with ground-state mass m. The energy difference between both configurations is given by√p 2 + M 2 − √ p 2 + m 2 ≈ (M 2 − m 2 )/2p. Therefore, the exp<strong>and</strong><strong>in</strong>g PLC can propagate for adistancel h ≃2pM 2 − m 2 , (4.11)the so-called hadronic expansion length or coherence length, before reach<strong>in</strong>g its normal hadronicsize.This length <strong>in</strong>creases with hadron momentum p <strong>and</strong> is <strong>in</strong>versely proportional to thesquared mass difference ∆M 2 between the <strong>in</strong>termediate PLC <strong>and</strong> the normal-sized hadron.The bare mass M of the PLC is an undeterm<strong>in</strong>ed parameter, but it is commonly assumed that0.7 ≤ ∆M 2 ≤ 1.1 (GeV/c 2 ) 2 are reasonable values.It is clear that the condition for full CT to occur, l h ≫ R A (with R A the nuclear radius), willnot be fulfilled for the present experimental k<strong>in</strong>ematics. The PLCs exp<strong>and</strong> <strong>and</strong> the IFSI are notcompletely suppressed. Consequently, the physical picture of the A(p, 2p) reaction with CT canbe described as follows: The imp<strong>in</strong>g<strong>in</strong>g proton compresses to a PLC as it hits a target nucleon,after which the outgo<strong>in</strong>g protons exp<strong>and</strong> from PLCs to normal-sized objects as they movethrough the nucleus. A schematic representation of this CT picture is given <strong>in</strong> Fig. 4.5. Here,it should be noted that PLCs are only formed <strong>in</strong> scatter<strong>in</strong>g via the quark-count<strong>in</strong>g mechanism,where the exchange of hard gluons (see Fig. 4.2) br<strong>in</strong>gs the quarks <strong>in</strong>side the hadrons closetogether. The L<strong>and</strong>shoff scatter<strong>in</strong>g does not possess this feature, s<strong>in</strong>ce the hard gluon exchangeis between the quarks of the different protons (see Fig. 4.4), which only makes the <strong>in</strong>teractionregion small.K 1P 1K 2Figure 4.5 Schematic representation of the CT phenomenon <strong>in</strong> the A(p, 2p) reaction. The <strong>in</strong>com<strong>in</strong>gproton is compressed to a PLC before hitt<strong>in</strong>g a target proton. The outgo<strong>in</strong>g protons are produced asPLCs but exp<strong>and</strong> as they move through the nucleus.


Chapter 4. <strong>Nuclear</strong> Transparency from A(p, 2p) Reactions 634.3.2 Quantum Diffusion Model of Farrar et al.To account for the reduced <strong>in</strong>teraction of a PLC with the nuclear medium, the total cross sectionsσpN tot <strong>in</strong> Eqs. (2.64) <strong>and</strong> (2.81) are replaced by effective ones.by [198], the effective cross section σpN efftive to the average size of a regular nucleon:σ effpN =As suggested for pQCDis simply scaled by the transverse size of the PLC rela-r2 t〈r 2 t 〉 σtot pN for r 2 t < 〈r 2 t 〉 . (4.12)Here, r t is the transverse radius of the PLC <strong>and</strong> 〈r 2 t 〉 1/2 is the rms transverse radius of a regularnucleon. It is assumed that the quarks <strong>in</strong> the PLC occupy a transverse area of r 2 t ≃ 〈n2 k 2 t 〉Q 2 〈r 2 t 〉at the po<strong>in</strong>t of <strong>in</strong>teraction, where Q 2 = |t| is the four-momentum transfer, n = 3 is the numberof constituents <strong>in</strong> the nucleon, <strong>and</strong> 〈k 2 t 〉 1/2 = 0.35 GeV/c is the average transverse momentumof a parton <strong>in</strong> a hadron. This assumption agrees with the Q 2 dependence of the transverse sizeof the nucleon found <strong>in</strong> the realistic models of the nucleon form factor [202].In the partonic model of Farrar et al. [148] (denoted by FLFS), two models are considered tof<strong>in</strong>d the dependence of r t on Z, the distance from the hard <strong>in</strong>teraction po<strong>in</strong>t along the trajectoryof the particle. In the naive quark expansion picture, the partons fly apart at the velocity of light,so that r t ∼ (E/m) −1 Z, with E/m the time dilatation factor. The hadronic expansion length isthen determ<strong>in</strong>ed by the Lorentz boost, l h ≃ (E/m)(σ totpN /π)1/2 . The alternate scenario followsfrom an analysis of pQCD Feynman diagrams [201,203], but this behavior is rather genericallycalled quantum diffusion. Consider<strong>in</strong>g the asymptotically most important energy denom<strong>in</strong>ator<strong>in</strong> these diagrams leads to r t ∼ Z 1/2 . The hadronic expansion length is determ<strong>in</strong>ed by theaverage value of the dom<strong>in</strong>ant energy denom<strong>in</strong>ator: l h ≃ 〈1/(E PLC − E)〉 ≃ 2p/∆M 2 (seeEq. (4.11)), where E PLC <strong>and</strong> E are the energy of the <strong>in</strong>termediate PLC <strong>and</strong> the normal-sizedhadron, respectively.Based on the above reason<strong>in</strong>g, the <strong>in</strong>teraction cross section is argued to be( {( Z) τσpNFLFS (p, Z) = σpNtot〈n 2 k +t 2 〉[ ( Z) τ ]})1 − θ(l h − Z) + θ(Z − l h ) , (4.13)l h |t| l hwhere τ = 1 corresponds to the quantum diffusion picture <strong>and</strong> τ = 2 to the naive quarkexpansion case. In this work, we shall present results only for the quantum diffusion model,τ = 1. The merit of this approach is its simplicity <strong>and</strong> physical motivations. It is, however,noth<strong>in</strong>g more than an educated guess, a semiclassical geometrical expansion model.4.3.3 Hadronic Expansion Model of Jenn<strong>in</strong>gs <strong>and</strong> MillerAn alternative perspective on CT was provided by the hadronic picture of Jenn<strong>in</strong>gs <strong>and</strong> Miller(JM) [149]. Their approach is completely quantum mechanical <strong>and</strong> avoids the use of semiclassicalapproximations. Us<strong>in</strong>g hadronic degrees of freedom, they suggested the follow<strong>in</strong>gexpression for the effective cross sectionσpN JM (p, Z) = σpNtot(1 − p p ∗ ei(p−p∗ )Z ) , (4.14)


4.3. Color Transparency 64with p the proton momentum <strong>and</strong> p ∗ the momentum of a baryon resonance with a complexmass M ∗ <strong>and</strong> the same energy as the nucleon, i.e., (p ∗ ) 2 = p 2 + M 2 p − (M ∗ ) 2 . The expression forthe effective cross section emanates from the <strong>in</strong>termediate PLC be<strong>in</strong>g a superposition of the nucleonground state <strong>and</strong> a nucleon resonance. The imag<strong>in</strong>ary part of M ∗ ensures the decay of the<strong>in</strong>termediate state to an asymptotically free, normal-sized proton. Later, Jenn<strong>in</strong>gs <strong>and</strong> Millerdeveloped a more sophisticated approach [204], where the <strong>in</strong>termediate PLC is exp<strong>and</strong>ed <strong>in</strong> acomplete set of hadronic states. This treatment uses experimentally measured matrix elementsfor deep <strong>in</strong>elastic scatter<strong>in</strong>g <strong>and</strong> diffractive dissociation to compute the effective cross section,thereby reduc<strong>in</strong>g the model dependence. In this work, however, we restrict ourselves to thesimple model of Eq. (4.14). Like the FLFS approach of Eq. (4.13), it considers one excited state<strong>in</strong> the PLC.It is worth not<strong>in</strong>g that both the FLFS <strong>and</strong> JM model take <strong>in</strong>to account the suppression of<strong>in</strong>teraction <strong>in</strong> the collision po<strong>in</strong>t <strong>and</strong> the time evolution of the PLC to a normal-sized protondur<strong>in</strong>g its propagation through the nucleus. This can be appreciated from the Z dependenceof the real part of the effective cross section shown <strong>in</strong> Fig. 4.6. The area below the curve is ameasure for the strength of the CT effect: the smaller this area, the more transparent the nuclearmedium. The Z = 0 <strong>in</strong>tercept shows that <strong>in</strong> the FLFS model the reduction <strong>in</strong> cross section is notcomplete at the collision po<strong>in</strong>t, σpNFLFS(Z= 0) goes like 1/Q2 . The FLFS effective cross sectionthen rises l<strong>in</strong>early up to σpN tot represent<strong>in</strong>g the expansion of the PLC. As mentioned before, theZ value at which this asymptotic value is reached, l h , depends on the proton momentum p <strong>and</strong>the parameter ∆M 2 . In the JM model, the real part of the effective cross section <strong>in</strong>creases fromzero at the <strong>in</strong>teraction po<strong>in</strong>t, then overshoots the total pN cross section <strong>and</strong> oscillates aboutthis value. This oscillat<strong>in</strong>g behavior dies out with <strong>in</strong>creas<strong>in</strong>g Z <strong>and</strong> Re σpN JM approaches σtotpNasymptotically. Likewise, the imag<strong>in</strong>ary part of σpN JM exhibits a damped oscillation around 0.The <strong>in</strong>itial <strong>in</strong>crease of Re σpN JM with Z is determ<strong>in</strong>ed by the real part of the parameter M ∗ : thelarger Re M ∗ , the sharper the <strong>in</strong>crease. The imag<strong>in</strong>ary part of M ∗ , on the other h<strong>and</strong>, has aneffect on the size of the oscillations of the effective cross section σpN JM around the asymptoticvalue, with small values for the imag<strong>in</strong>ary part lead<strong>in</strong>g to large oscillations.4.3.4 Effect of CT on the IFSI FactorIn Fig. 4.7 the absolute value of the IFSI factor is shown for two values of the <strong>in</strong>com<strong>in</strong>g beammomentum p 1 as a function of the (x, z) coord<strong>in</strong>ate of the hard collision po<strong>in</strong>t. The absolutevalue of the IFSI factor is a measure for the total absorption due to the <strong>in</strong>com<strong>in</strong>g <strong>and</strong> outgo<strong>in</strong>gprotons. The k<strong>in</strong>ematics is chosen such that the c.m. scatter<strong>in</strong>g angle for the elementaryhard pp scatter<strong>in</strong>g is 90 ◦ . Accord<strong>in</strong>gly, the pair of f<strong>in</strong>al-state protons is produced at equal momenta,equal polar angles, <strong>and</strong> opposite azimuthal angles <strong>in</strong> the laboratory frame. The scatter<strong>in</strong>gangles of the two f<strong>in</strong>al-state protons become smaller as the <strong>in</strong>cident beam energy <strong>in</strong>creasesfor a fixed c.m. scatter<strong>in</strong>g angle. At p 1 = 5.9 GeV/c, the outgo<strong>in</strong>g protons have momentak 1 = k 2 = 3.3 GeV/c <strong>and</strong> scatter<strong>in</strong>g angles θ 1 = θ 2 = 27.5 ◦ ; while at p 1 = 14.4 GeV/c, the


Chapter 4. <strong>Nuclear</strong> Transparency from A(p, 2p) Reactions 651FLFS model1.81.6JM modeleff totRe ( σ pN/ σ pN)0.80.60.40.2∆M 2 =1.1 (GeV/c 2 ) 2∆M 2 =0.7 (GeV/c 2 ) 200 1 2 3 4 5 6 7 8 9 10Z (fm)tot)Re ( σ pNeff/ σ pN1.41.210.80.6M * = (1770 + 75 i) MeV0.4M * = (1440 + 75 i) MeV0.2M * = (1770 + 150 i) MeVM * = (1440 + 150 i) MeV00 1 2 3 4 5 6 7 8 9 10Z (fm)Figure 4.6 The real part of σpN eff /σtot pN as a function of the distance Z from the hard <strong>in</strong>teraction po<strong>in</strong>t atp ≃ 5.4 GeV/c <strong>and</strong> |t| ≃ 8.5 (GeV/c) 2 . The left panel shows the Z dependence of the effective crosssection <strong>in</strong> the FLFS model, while the right panel displays the results of the JM model.k<strong>in</strong>ematics is k 1 = k 2 = 7.6 GeV/c <strong>and</strong> θ 1 = θ 2 = 19.3 ◦ . The st<strong>and</strong>ard RMSGA results arecompared with the RMSGA + CT predictions of the FLFS model us<strong>in</strong>g ∆M 2 = 0.7 (GeV/c 2 ) 2 .The RMSGA results show that the effect of IFSI is mostly constra<strong>in</strong>ed to low impact parameter|x|, i.e., trajectories that pass through the dense nuclear <strong>in</strong>terior. The overall behavior ofthe IFSI factor <strong>in</strong> the RMSGA approach is very similar for both <strong>in</strong>com<strong>in</strong>g momenta. After all,the attenuation effect is controlled by the total pN cross section σpN tot which, at the high energies<strong>in</strong> question, is almost constant (see Fig. 2.4). The ma<strong>in</strong> difference between both k<strong>in</strong>ematics is<strong>in</strong> the scatter<strong>in</strong>g angles θ 1 = θ 2 , i.e., it is purely geometrical. This causes the dissimilaritiesbetween the p 1 = 5.9 GeV/c <strong>and</strong> p 1 = 14.4 GeV/c IFSI factors <strong>in</strong> the backward hemisphere(z < 0), where the FSI effects are largest. Turn<strong>in</strong>g to CT, its effect is particularly prom<strong>in</strong>entfor hard <strong>in</strong>teraction po<strong>in</strong>ts that are located near the center of the nucleus. For these po<strong>in</strong>ts, theregion where the protons are <strong>in</strong> PLCs (i.e., where the <strong>in</strong>teraction with the nuclear medium ismost reduced) co<strong>in</strong>cides with the part of the nucleus where the spectator nucleons are mostabundant.4.3.5 <strong>Nuclear</strong> Color Screen<strong>in</strong>g EffectTo complete the discussion on CT, we mention the nuclear color screen<strong>in</strong>g effect (NCSE) [203,205]. In this QCD effect, the b<strong>in</strong>d<strong>in</strong>g of the nucleonic system results <strong>in</strong> a reduced PLC probability.S<strong>in</strong>ce the potential for the <strong>in</strong>teraction of a bound nucleon <strong>in</strong> a PLC with nearby nucleonsis smaller than for a normal-sized nucleon, the creation of a PLC gives rise to smaller b<strong>in</strong>d<strong>in</strong>genergies. This is not preferable energywise <strong>and</strong> leads to the suppression of the nucleon’s PLCcomponent. This suppression can be <strong>in</strong>cluded <strong>in</strong> the calcutions by multiply<strong>in</strong>g the CT cross


4.4. <strong>Nuclear</strong> Transparency Results 66RMSGA, p 1 = 5.9 GeV/cRMSGA+CT, p 1 = 5.9 GeV/c|S IFSI|0.80.60.40.2|S IFSI|0.80.60.40.2642x (fm)-6 -4 -2 0 20-2-4-64z (fm)6642x (fm)-6 -4 -2 0 20-2-4-64z (fm)6RMSGA, p 1 = 14.4 GeV/cRMSGA+CT, p 1 = 14.4 GeV/c|S IFSI|0.80.60.40.2|S IFSI|0.80.60.40.2642x (fm)-6 -4 -2 0 20-2-4-64z (fm)6642x (fm)-6 -4 -2 0 20-2-4-64z (fm)6Figure 4.7 The (x, z) dependence of the absolute value of the IFSI factor S IFSI <strong>in</strong> the RMSGA (left panels)<strong>and</strong> RMSGA + CT picture (right panels). (x, z) is the coord<strong>in</strong>ate of the hard collision po<strong>in</strong>t <strong>in</strong> the scatter<strong>in</strong>gplane with the z axis ly<strong>in</strong>g along the <strong>in</strong>com<strong>in</strong>g momentum ⃗p 1 . In the CT calculations, the FLFSmodel of Eq. (4.13) with ∆M 2 = 0.7 (GeV/c 2 ) 2 was employed. The upper (lower) panels correspond toan <strong>in</strong>com<strong>in</strong>g momentum of 5.9 GeV/c (14.4 GeV/c). The k<strong>in</strong>ematics of the outgo<strong>in</strong>g protons is <strong>in</strong>-plane<strong>and</strong> symmetric, def<strong>in</strong>ed by 90 ◦ pp scatter<strong>in</strong>g <strong>in</strong> the pp c.m. frame.section by the factor[ (δk = θ(Q 2 0 − Q 2 ) + θ(Q 2 − Q 2 0) 1 + 1 − Q2 ) k20 M p+ 2ɛ] −2 AQ 2 . (4.15)∆EHere, Q 2 = |t| is the four-momentum transfer, k is the bound nucleon’s momentum, <strong>and</strong> ɛ A ≃8 MeV is the average b<strong>in</strong>d<strong>in</strong>g energy per nucleon. For the parameter ∆E, we will take a valueof 0.6 GeV, while an analysis of the 2 He(e, e ′ ) SLAC data <strong>in</strong> Ref. [203] <strong>in</strong>dicates that Q 2 02 (GeV/c) 2 .4.4 <strong>Nuclear</strong> Transparency ResultsExperimentally, the nuclear transparency is def<strong>in</strong>ed as the ratio of the cross section for quasielasticscatter<strong>in</strong>g from the protons <strong>in</strong> the nucleus to the cross section for free pp scatter<strong>in</strong>g corrected≃


Chapter 4. <strong>Nuclear</strong> Transparency from A(p, 2p) Reactions 67for the number of protons <strong>in</strong> the nucleus Z:T =σ (pp quasielastic <strong>in</strong> nucleus)Z σ (pp elastic <strong>in</strong> hydrogen) . (4.16)In our calculations, the nuclear transparency is computed as the ratio of the A(p, 2p) crosssections <strong>in</strong>clud<strong>in</strong>g <strong>and</strong> exclud<strong>in</strong>g IFSI effects:T =σ(p,2p)σ (p,2p)RPWIA. (4.17)The RPWIA limit is reached by sett<strong>in</strong>g the IFSI operator ŜRMSGA(+CT)IFSI(⃗r) equal to one <strong>in</strong>Eq. (4.10) <strong>and</strong> as such the free pp cross section is recovered <strong>in</strong> the denom<strong>in</strong>ator. The numerator<strong>and</strong> denom<strong>in</strong>ator of Eq. (4.17) are obta<strong>in</strong>ed by <strong>in</strong>tegrat<strong>in</strong>g the correspond<strong>in</strong>g differentialcross sections over the phase space def<strong>in</strong>ed by the k<strong>in</strong>ematic cuts. In our calculations, weadopted identical cuts as <strong>in</strong> the experiments [124–126] <strong>and</strong> assumed a flat experimental acceptancewith<strong>in</strong> the k<strong>in</strong>ematical ranges for each data po<strong>in</strong>t. These experimental cuts, which areapplied to extract the quasielastic events from the background, constra<strong>in</strong> the values of p mx ,p my , α 0 , <strong>and</strong> θ c.m. . Here, the z direction is def<strong>in</strong>ed to co<strong>in</strong>cide with the <strong>in</strong>cident beam direction,so p mx <strong>and</strong> p my are the transverse components of the miss<strong>in</strong>g momentum. The variable√( ) ( )(E p1 + M p ) 2 − 4Mp 2 cos θ1 −θ 22cos θ1 +θ 22− p 1α 0 ≡ 1 −, (4.18)M pis an approximation to α ≡ A Em−pmzM A, the longitud<strong>in</strong>al light-cone momentum fraction carriedby the struck proton [206]. The range of the c.m. scatter<strong>in</strong>g angle extends from 80 ◦ to 90 ◦for the E834 experiment by Carroll et al. [124], while the sequel E850 experiment [125, 126]covers the region 86 ◦ θ c.m. 90 ◦ . The c.m. scatter<strong>in</strong>g angle can be determ<strong>in</strong>ed directly bytransform<strong>in</strong>g the outgo<strong>in</strong>g particle momentum to the c.m. frame <strong>and</strong> calculat<strong>in</strong>g its dot productwith the beam momentum. Unfortunately, this procedure generates a result with unsatisfactoryresolution. Therefore, θ c.m. is calculated us<strong>in</strong>g the approximate relation [206]( ) ( )2p 1√(E p1 + M p ) 2 − 4Mp 2 s<strong>in</strong> θ1 −θ 22s<strong>in</strong> θ1 +θ 22cos θ c.m. =s − 4Mp2 . (4.19)For fixed beam energy E p1 , α 0 <strong>and</strong> θ c.m. depend ma<strong>in</strong>ly on the sum <strong>and</strong> difference of the polarangles θ 1 <strong>and</strong> θ 2 of the outgo<strong>in</strong>g nucleons, respectively. The <strong>in</strong>troduction of the approximations(4.18) <strong>and</strong> (4.19) turned out to be essential to <strong>in</strong>terpret the experiments, as they significantlyimproved the resolution.We consider the experimental nuclear transparency values as presented <strong>in</strong> [207]. The <strong>in</strong>cidentlab momentum varies from 5.9 to 14.4 GeV/c <strong>and</strong> the scatter<strong>in</strong>g angle is near 90 ◦ <strong>in</strong> the ppcenter of mass. The M<strong>and</strong>elstam variable |t| ≃ ( s − 4M 2 p)/2 extends from 4.7 to 12.7 (GeV/c) 2 .In Ref. [207], the orig<strong>in</strong>ally published values of Carroll et al. [124] were rescaled us<strong>in</strong>g the improvednuclear momentum distributions of Ref. [208], thereby mak<strong>in</strong>g them consistent withthe data of Refs. [125, 126].


4.4. <strong>Nuclear</strong> Transparency Results 68First, we address the energy dependence of the 12 C(p, 2p) transparency <strong>and</strong> study the roleof IFSI, CT, <strong>and</strong> NF. Further, we use the 12 C(p, 2p) calculations as a test case to determ<strong>in</strong>e the accuracyof the IFSI operator of Eq. (2.81) relative to the expression of Eq. (2.74). Fig. 4.8 displaysthe 12 C transparency as a function of the <strong>in</strong>com<strong>in</strong>g proton momentum p 1 . The solid curves representthe full RMSGA calculations, whereas the RMSGA ′ results are shown as dashed curves.Three different scenarios were considered. As expected, the st<strong>and</strong>ard RMSGA calculations leadto a nuclear transparency that is almost <strong>in</strong>dependent of the beam momentum. The ma<strong>in</strong> effectof the IFSI is to reduce the nuclear transparency from the asymptotic value of 1 to ∼ 0.15.The <strong>in</strong>clusion of CT effects produces a transparency l<strong>in</strong>early ris<strong>in</strong>g with energy. The <strong>in</strong>creaserelative to the RMSGA result is highly dependent on the adopted model <strong>and</strong> correspond<strong>in</strong>gparameters for CT. The curves <strong>in</strong>clud<strong>in</strong>g CT shown <strong>in</strong> Fig. 4.8 adopt the FLFS model with∆M 2 = 0.7 (GeV/c 2 ) 2 . The <strong>in</strong>crease of the transparency is consistent with the data <strong>in</strong> therange 5–10 GeV/c, but the RMSGA + CT picture fails to expla<strong>in</strong> the drop <strong>in</strong> the transparencyat higher momenta. Our RMSGA <strong>and</strong> RMSGA + CT predictions confirm the results of [209].A better agreement with the data is obta<strong>in</strong>ed when add<strong>in</strong>g the mechanism of NF. Compared tothe RMSGA + CT results, the transparency is <strong>in</strong>creased at <strong>in</strong>termediate momenta (5–10 GeV/c)<strong>and</strong> decreased at higher momenta, two effects which improve the description of the data. Asimilar result was obta<strong>in</strong>ed <strong>in</strong> Ref. [210] where the JM model of CT was used.0.60.50.4RMSGARMSGA’RMSGARMSGA+CTRMSGA+CT+NFCarrollMardorLeksanovT0.30.20.105 6 7 8 9 10 11 12 13 14 15p 1(GeV/c)Figure 4.8 The nuclear transparency for the 12 C(p, 2p) reaction as a function of the <strong>in</strong>com<strong>in</strong>g lab momentump 1 . The full RMSGA (solid l<strong>in</strong>es) are compared to the RMSGA ′ (dashed l<strong>in</strong>es) results. The differentcurves represent the RMSGA, RMSGA + CT <strong>and</strong> RMSGA + CT + NF calculations. The CT effects arecalculated <strong>in</strong> the FLFS model [148] with ∆M 2 = 0.7 (GeV/c 2 ) 2 <strong>and</strong> the results <strong>in</strong>clud<strong>in</strong>g the mechanismof NF are obta<strong>in</strong>ed us<strong>in</strong>g the positive sign of φ (s) + δ 1 . Data are from Refs. [124] (solid triangles), [125](solid squares), <strong>and</strong> [126] (solid circles).


Chapter 4. <strong>Nuclear</strong> Transparency from A(p, 2p) Reactions 69Concern<strong>in</strong>g the comparison of the RMSGA <strong>and</strong> RMSGA ′ results, it can be <strong>in</strong>ferred fromFig. 4.8 that both approaches yield nearly identical results which differ at the 2–3% level. Consequently,the operator of Eq. (2.81) is considered sufficiently accurate for the calculation of thenuclear transparency <strong>and</strong> will be used <strong>in</strong> the rema<strong>in</strong>der of this work. We wish to stress thatthe computational cost of Eq. (2.81) is about a factor of 10 3 lower than the full-blown RMSGAoperator of Eq. (2.74).Figs. 4.9 <strong>and</strong> 4.10 are devoted to a comparison of the different CT models. Results of theFLFS quantum diffusion model are plotted for ∆M 2 = 0.7 <strong>and</strong> 1.1 (GeV/c 2 ) 2 . For the M ∗parameter of the JM model we consider three different values, represent<strong>in</strong>g the ∆, the Roperresonance, <strong>and</strong> the average of the lowest P -wave N ∗ resonances. For the imag<strong>in</strong>ary part of M ∗a value of 150 MeV was taken. The lowest values of the parameters ∆M 2 <strong>and</strong> M ∗ <strong>in</strong>duce thestrongest <strong>in</strong>crease of the RMSGA + CT transparency with the beam momentum p 1 <strong>and</strong> alsolead to the largest deviations between the predictions <strong>in</strong>clud<strong>in</strong>g the NF mechanism <strong>and</strong> thecorrespond<strong>in</strong>g RMSGA + CT results.Fig. 4.9 shows that after <strong>in</strong>clud<strong>in</strong>g NF <strong>and</strong> CT, the calculations correctly reproduce the maximum<strong>in</strong> the 12 C(p, 2p) transparency at about 9.5 GeV/c, but badly fail to fall back low enoughto account for the 14.4 GeV/c data po<strong>in</strong>t. Ralston <strong>and</strong> Pire [144], on the other h<strong>and</strong>, do succeed<strong>in</strong> reproduc<strong>in</strong>g the maxima <strong>and</strong> m<strong>in</strong>ima <strong>in</strong> the nuclear transparency data. However, they assumethat the transparency of the quark-count<strong>in</strong>g term is beam-energy <strong>in</strong>dependent, a ratherspeculative assumption. The CT-<strong>in</strong>duced <strong>in</strong>crease of the quark-count<strong>in</strong>g transparency with energycauses our RMSGA + CT + NF predictions to rise aga<strong>in</strong> at a momentum p 1 ≃ 12 GeV/c.In the FLFS as well as the JM approach, the RMSGA + CT + NF predictions reproducethe general trend of the data, but no variant achieves very good agreement. Furthermore, itis not possible to unambiguously determ<strong>in</strong>e the value of the parameters ∆M 2 <strong>and</strong> M ∗ , as thebest choice for these parameters depends on the target nucleus under consideration. Fig. 4.9suggests that for the 12 C(p, 2p) reaction M ∗ = (1440 + 150 i) MeV leads to the best agreement,while for the ∆M 2 parameter no “best” choice can be put forward. Fig. 4.10, on the other h<strong>and</strong>,shows that for the 27 Al target nucleus the FLFS results are systematically below the data <strong>in</strong>the region below 10 GeV/c. Us<strong>in</strong>g M ∗ = (1440 + 150 i) MeV <strong>in</strong> the JM model also does not<strong>in</strong>crease the transparency high enough so as to match the 6 <strong>and</strong> 10 GeV/c data po<strong>in</strong>ts, onlywith M ∗ = (1232 + 150 i) MeV is the CT-<strong>in</strong>duced <strong>in</strong>crease of the transparency strong enough.Another effect that can be studied <strong>in</strong> Figs. 4.9 <strong>and</strong> 4.10 is the <strong>in</strong>fluence of the sign of φ (s) +δ 1 on the RMSGA + CT + NF results. For the FLFS model of CT, the differences betweencalculations us<strong>in</strong>g both signs of φ (s)+δ 1 are m<strong>in</strong>or. As already observed by Jenn<strong>in</strong>gs <strong>and</strong> Miller[210], the results us<strong>in</strong>g the JM model of CT are rather sensitive to this sign. The discrepancybetween the FLFS- <strong>and</strong> JM-based calculations arises from the different structure of the effectivecross sections (4.13) <strong>and</strong> (4.14). Indeed, these effective cross sections not only determ<strong>in</strong>e theattenuation of the quark-count<strong>in</strong>g term <strong>in</strong> the nuclear medium, but also the phase differencebetween the quark-count<strong>in</strong>g <strong>and</strong> the L<strong>and</strong>shoff term. Whereas the real parts of both effective


4.4. <strong>Nuclear</strong> Transparency Results 700.60.5∆M 2 = 0.7 (GeV/c 2 ) 2∆M 2 = 1.1 (GeV/c 2 ) 20.40.30.2T0.10.6positive sign φ(s)+δ 1negative sign φ(s)+δ 1without L<strong>and</strong>shoff contribution0.50.40.30.20.1 M * = (1232 + 150 i) MeVM * = (1440 + 150 i) MeVM * = (1770 + 150 i) MeV05 6 7 8 9 10 11 12 13 14 15p 1(GeV/c)Figure 4.9 The 12 C(p, 2p) transparency versus the <strong>in</strong>com<strong>in</strong>g lab momentum p 1 . The upper (lower) paneldepicts results us<strong>in</strong>g the FLFS (JM) model for CT. Calculations <strong>in</strong>clud<strong>in</strong>g the effects of CT <strong>and</strong> NF withthe positive (solid l<strong>in</strong>es) <strong>and</strong> negative (dashed l<strong>in</strong>es) sign for φ (s)+δ 1 are shown, along with the RMSGA+ CT predictions (dot-dashed l<strong>in</strong>es). Data are from Refs. [124] (solid triangles), [125] (solid squares),<strong>and</strong> [126] (solid circles).cross sections are quite similar (see Fig. 4.6), the FLFS effective cross section (4.13) is purelyreal, while its JM counterpart has an imag<strong>in</strong>ary part as well. This imag<strong>in</strong>ary part causes theenhanced sensitivity of the JM results to the sign of φ (s) + δ 1 . As for which sign causes the bestagreement with the data, no firm conclusions can be drawn. Indeed, the optimum choice forthe sign of φ (s) + δ 1 depends on the used CT model (FLFS or JM), the value of the parameter∆M 2 or M ∗ , <strong>and</strong> the target nucleus. For the 7 Li(p, 2p) reaction, the positive sign provides thebest agreement, while the 63 Cu transparency data rather require a negative sign.None of the results shown <strong>in</strong> this section <strong>in</strong>clude the nuclear color screen<strong>in</strong>g effect of Section4.3.5. We deem that the CT parameters are so badly constra<strong>in</strong>ed that controll<strong>in</strong>g additionalmechanisms is out of reach for the moment. In both the RMSGA + CT <strong>and</strong> the RMSGA + CT+ NF calculations, the <strong>in</strong>clusion of the NCSE decreases the transparency by 6–12%, with thelargest effect occur<strong>in</strong>g at higher p 1 . Another additional effect that is not <strong>in</strong>cluded here are the


Chapter 4. <strong>Nuclear</strong> Transparency from A(p, 2p) Reactions 710.30.25∆M 2 = 0.7 (GeV/c 2 ) 2∆M 2 = 1.1 (GeV/c 2 ) 20.20.150.1T0.050.3positive sign φ(s)+δ 1negative sign φ(s)+δ 1without L<strong>and</strong>shoff contribution0.250.20.150.10.050M * = (1232 + 150 i) MeVM * = (1440 + 150 i) MeVM * = (1770 + 150 i) MeV6 7 8 9 10 11 12p 1(GeV/c)Figure 4.10 As <strong>in</strong> Fig. 4.9, but for 27 Al. Data are from Ref. [124].NN short-range correlations (SRC). It should be noted that the SRC <strong>and</strong> CT both lead to asuppression of the absorptive <strong>in</strong>teraction close to the hard collision po<strong>in</strong>t <strong>and</strong> therefore are <strong>in</strong>some sense compet<strong>in</strong>g mechanisms. However, the effects stemm<strong>in</strong>g from SRC are <strong>in</strong>dependentof energy, <strong>and</strong>, hence, can be disentangled from the pure CT effects. The SRC were found to<strong>in</strong>crease the st<strong>and</strong>ard distorted-wave A(p, 2p) transparency (i.e., without CT or NF taken <strong>in</strong>toaccount) by about 30%, without chang<strong>in</strong>g the energy dependence [209, 211].Study<strong>in</strong>g the A dependence of the experimental nuclear transparency is a very efficient toolto search for CT effects. Indeed, a change <strong>in</strong> the shape of the A dependence with <strong>in</strong>creas<strong>in</strong>g<strong>in</strong>com<strong>in</strong>g momentum could <strong>in</strong>dicate the onset of CT, with complete CT correspond<strong>in</strong>g to avanish<strong>in</strong>g A dependence. The A dependence of the nuclear transparency at two values of the<strong>in</strong>com<strong>in</strong>g momentum p 1 is studied <strong>in</strong> Fig. 4.11. The st<strong>and</strong>ard RMSGA calculations fall considerablybelow the data. Further, none of the RMSGA + CT + NF calculations succeed <strong>in</strong>simultaneously describ<strong>in</strong>g the data for all target nuclei. While the FLFS approach agrees with


4.5. Outlook 72the 7 Li <strong>and</strong> 12 C data po<strong>in</strong>ts rather well us<strong>in</strong>g ∆M 2 = 0.7 (GeV/c 2 ) 2 , the FLFS results tend tounderestimate the 27 Al <strong>and</strong> 63 Cu data. With regard to the M ∗ parameter of the JM model, forthe 7 Li <strong>and</strong> 12 C nuclei a value of M ∗ = (1440 + 150 i) MeV seems acceptable, whereas the heavier27 Al <strong>and</strong> 63 Cu nuclei need a smaller M ∗ value. A general feature of the RMSGA + CT +NF predictions is that their A dependence is steeper than the data. The RMSGA calculations ofthe A(e, e ′ p) transparency by our research group [87] also display this characteristic of overestimat<strong>in</strong>gthe measured A dependence. F<strong>in</strong>ally, the dot-dashed <strong>and</strong> dotted curves <strong>in</strong>dicate thatthe 6 GeV/c data are proportional to A −2/3 ; while at 10 GeV/c, the A dependence of the data ismore gradual (T ∝ A −1/3 ). This trend could po<strong>in</strong>t to the onset of CT <strong>and</strong> is not reproduced bythe st<strong>and</strong>ard RMSGA predictions, which are almost <strong>in</strong>dependent of the <strong>in</strong>com<strong>in</strong>g momentum.When CT <strong>and</strong> NF effects are <strong>in</strong>cluded, the soften<strong>in</strong>g of the A dependence with <strong>in</strong>creas<strong>in</strong>g <strong>in</strong>com<strong>in</strong>gmomentum is also present <strong>in</strong> the calculations, albeit not as pronounced as <strong>in</strong> the data.For example, the FLFS calculations with ∆M 2 = 0.7 (GeV/c 2 ) 2 correspond to T ∝ A −0.83 at6 GeV/c <strong>in</strong>com<strong>in</strong>g momentum <strong>and</strong> T ∝ A −0.70 at 10 GeV/c.4.5 OutlookTo conclude this chapter, we present some clos<strong>in</strong>g remarks <strong>and</strong> po<strong>in</strong>t out some suggestions forfurther research.Introduc<strong>in</strong>g the concept of nuclear filter<strong>in</strong>g, our calculations can be brought <strong>in</strong> qualitativeagreement with the A(p, 2p) transparency data, thereby confirm<strong>in</strong>g earlier calculations[204, 210, 212]. We wish to stress that CT is imperative to <strong>in</strong>crease the predicted transparenciesto the level of the data. A similar conclusion was drawn <strong>in</strong> Refs. [212, 213]. The quantitativedescription of the data, however, is far from perfect <strong>and</strong> it is not possible to constra<strong>in</strong> the magnitudeof the parameters <strong>in</strong> the CT models. More specifically, the oscillations seem to be moreoutspoken <strong>in</strong> the data than <strong>in</strong> the calculations. This is because <strong>in</strong>clud<strong>in</strong>g the non-zero absorptionof the exp<strong>and</strong><strong>in</strong>g PLC <strong>and</strong> the non-complete absorption of the normal-sized proton tendsto make the two terms similar <strong>and</strong> weakens the <strong>in</strong>terference effects.The <strong>in</strong>dications for CT <strong>in</strong> A(p, 2p) reactions are not necessarily <strong>in</strong> contradiction with theresults from A(e, e ′ p) experiments. Although the A(e, e ′ p) transparencies exhibit no significant<strong>in</strong>crease with the four-momentum transfer Q 2 [214–218] <strong>and</strong> can be reasonably reproduced <strong>in</strong>the RMSGA framework [87], the existence of CT can not be excluded s<strong>in</strong>ce the predicted effectis small [197, 219–221]. The calculations of Lava [221] are depicted <strong>in</strong> Fig. 4.12 <strong>and</strong> show thatthere is <strong>in</strong>deed no conclusive evidence for the onset of CT <strong>in</strong> A(e, e ′ p) reactions. The smallnessof the CT effect <strong>in</strong> the A(e, e ′ p) experiments is caused by the short expansion times of the PLCto normal size at the present k<strong>in</strong>ematics. The effect of CT is more pronounced <strong>in</strong> the A(p, 2p)transparency for different reasons. First, <strong>in</strong> the A(p, 2p) reaction there are three particles thatcan experience CT <strong>in</strong>stead of only one <strong>in</strong> A(e, e ′ p) reactions. Second, A(p, 2p) data are availableup to Q 2 = |t| values of 12.7 (GeV/c) 2 , while A(e, e ′ p) transparency experiments are restrictedto Q 2 8 (GeV/c) 2 .


Chapter 4. <strong>Nuclear</strong> Transparency from A(p, 2p) Reactions 730.60.50.40.3RMSGA’FLFS: ∆M 2 = 0.7 (GeV/c 2 ) 2FLFS: ∆M 2 = 1.1 (GeV/c 2 ) 2JM: M * = (1232 + 150 i) MeVJM: M * = (1440 + 150 i) MeVT ∝ A -1/3T ∝ A -2/30.2T0.10.6p 1= 6 GeV/c0.50.40.30.20.10p 1= 10 GeV/c10AFigure 4.11 A dependence of the nuclear transparency at two values of the <strong>in</strong>com<strong>in</strong>g lab momentum p 1 .The st<strong>and</strong>ard RMSGA calculations are represented by dashed curves, while the solid curves are RMSGA+ CT + NF calculations with the positive sign for φ (s) + δ 1 . The solid curves correspond with differentdescriptions of the CT effects, as <strong>in</strong>dicated by the legend. The dot-dashed (dotted) curves display theA −1/3 (A −2/3 ) parametrization, normalized to the 7 Li data po<strong>in</strong>ts. Data are from Ref. [124].Despite the fact that CT is expected to reveal itself more rapidly <strong>in</strong> the A(p, 2p) reaction,there are some disadvantages. The free pp scatter<strong>in</strong>g data show oscillations which are ratherclear evidence for quantum mechanical <strong>in</strong>terference of two amplitudes. For the pp scatter<strong>in</strong>g<strong>in</strong>side the nucleus, however, the oscillations are much reduced. An <strong>in</strong>dication for this is thecorrelation between the <strong>in</strong>creases <strong>in</strong> the transparency ratio <strong>and</strong> the decreases <strong>in</strong> the free-spacedata, <strong>and</strong> vice versa. Furthermore, a plot of the nuclear differential cross section multipliedby the overall power-law factor s 10 shows a reasonably flat s dependence (see Ref. [224] <strong>and</strong>Fig. 4.13). These observations support the idea that pp scatter<strong>in</strong>g <strong>in</strong>side the nuclear mediumdiffers from free pp scatter<strong>in</strong>g. Hence, the experimentally attractive transparency ratio is not areliable observable to measure CT on its own. Indeed, the experimental nuclear transparencyis not a pure survival probability, but also reflects the effect of the nuclear medium on the


4.5. Outlook 7410.812 C0.6Transparency0.40.210.8∆M 2 = 1.1 (GeV/c 2 ) 2∆M 2 = 0.7 (GeV/c 2 ) 2RMSGA56 Fe0.60.40.200 1 2 3 4 5 6 7 8 9 10Q 2 (GeV 2 /c 2 )Figure 4.12 The A(e, e ′ p) nuclear transparency versus Q 2 <strong>in</strong> quasielastic k<strong>in</strong>ematics. The upper (lower)panel displays the results for 12 C ( 56 Fe). The solid l<strong>in</strong>e shows the st<strong>and</strong>ard RMSGA results. The dashed(dot-dashed) curves account for the effect of CT <strong>in</strong> the FLFS model with ∆M 2 = 0.7 (1.1) (GeV/c 2 ) 2 .Data are from Refs. [222] (open squares), [214, 215] (open triangles), [217] (solid circles), [216, 218] (solidtriangles), <strong>and</strong> [223] (open diamonds). This figure is taken from Ref. [221].underly<strong>in</strong>g hard wide-angle scatter<strong>in</strong>g. In other words, the NF effect obscures the observationof CT.A number of uncerta<strong>in</strong>ties <strong>in</strong>volv<strong>in</strong>g the A(p, 2p) transparency rema<strong>in</strong>. One subject of discussionis the size of the L<strong>and</strong>shoff term. Accord<strong>in</strong>g to Botts et al. [185, 186, 225–227], this termmight also be small-sized. This would br<strong>in</strong>g the survival probability of the L<strong>and</strong>shoff <strong>and</strong> thequark-count<strong>in</strong>g term closer together <strong>and</strong> would further weaken the oscillations <strong>in</strong> the energydependence of the computed transparencies, an effect that the data do not seem to support.It might also be of <strong>in</strong>terest to exam<strong>in</strong>e the <strong>in</strong>fluence of the helicity-nonconserv<strong>in</strong>g amplitudeson the A(p, 2p) transparency. Recently, it was shown that <strong>in</strong>clud<strong>in</strong>g the helicity-nonconserv<strong>in</strong>gamplitudes <strong>and</strong> their <strong>in</strong>terference with the L<strong>and</strong>shoff amplitude leads to a better description ofthe pp elastic scatter<strong>in</strong>g cross section <strong>and</strong> sp<strong>in</strong> correlation [228].As mentioned before (see Section 4.1.4), apart from the Ralston-Pire picture on which ourresults are based, other explanations of the energy dependence of the transparency have beensuggested. At the present energies, both the Ralston-Pire approach <strong>and</strong> the dibaryon reso-


Chapter 4. <strong>Nuclear</strong> Transparency from A(p, 2p) Reactions 75Figure 4.13 The nuclear cross section T dσ/dt multiplied by an overall power-law factor s 10 . This plotis taken from Ref. [224] <strong>and</strong> shows that the oscillations <strong>in</strong> the nuclear cross section are fairly small.nance model of Brodsky <strong>and</strong> de Teramond can accomodate the experimental transparencies.At higher energies, however, the predictions of both models are very different. An improvedset of data, particularly at higher energies, is essential to dist<strong>in</strong>guish between both models.The 50 GeV proton synchrotron that is under construction at J-PARC [229, 230] opens greatopportunities for this research.F<strong>in</strong>ally, the θ c.m. dependence of the nuclear transparency is as yet unexpla<strong>in</strong>ed. At an <strong>in</strong>com<strong>in</strong>gmomentum of 5.9 GeV/c, the data show a significant dependence on the c.m. scatter<strong>in</strong>gangle, while at 7.5 GeV/c the θ c.m. dependence flattens out [125, 206]. Most theoretical predictions,<strong>in</strong>clud<strong>in</strong>g ours, do not heavily depend on θ c.m. . The θ c.m. dependence is similar only tothe <strong>in</strong>verse of the sp<strong>in</strong> correlation, (A NN ) −1 , <strong>and</strong> this possible relation between A NN <strong>and</strong> nucleartransparency is certa<strong>in</strong>ly worthy of further <strong>in</strong>vestigation. In that respect, the possibilityof A(p, 2p) transparency measurements at θ c.m. = 15–40 ◦ <strong>in</strong> the same Q 2 region (Q 2 = 2–10 (GeV/c) 2 ) at the Serpukhov 70 GeV accelerator [231] sounds very promis<strong>in</strong>g.


4.5. Outlook 76


Chapter 5Second-Order Eikonal Corrections <strong>in</strong>A(e, e ′ p) ObservablesThe previous chapter dealt with the issue of scatter<strong>in</strong>g <strong>in</strong> the GeV energy regime <strong>and</strong> the accompany<strong>in</strong>gphenomena like CT. In this chapter, we will turn our attention to lower energies<strong>and</strong> <strong>in</strong>vestigate the low-energy accuracy of the eikonal approach. Hereby, our focus is on exclusiveelectro-<strong>in</strong>duced A(e, e ′ p) reactions. A(e, e ′ p) processes provide access to a wide rangeof nuclear phenomena like short- <strong>and</strong> long-range correlations, relativistic effects, the transitionfrom hadronic to partonic degrees of freedom, <strong>and</strong> medium modifications of nucleon properties.Just like A(p, pN) reactions, the <strong>in</strong>terpretation of A(e, e ′ p) data heavily relies on an accuratedescription of the effect of the FSI, i.e., the <strong>in</strong>teractions of the ejected proton with the residualnucleus such as rescatter<strong>in</strong>g <strong>and</strong>/or absorption. The eikonal approximation has been widelyused to treat these distortions, either <strong>in</strong> conjunction with optical potentials [56, 57, 72–75], orwith Glauber theory, its multiple-scatter<strong>in</strong>g extension [57, 59, 80–86]. The eikonal approach,however, is commonly used <strong>in</strong> a first-order approximation. The purpose of this chapter is todeterm<strong>in</strong>e the <strong>in</strong>fluence of second-order eikonal corrections on A(e, e ′ p) observables by us<strong>in</strong>gthe SOROMEA model that was developed <strong>in</strong> Section 2.5.The outl<strong>in</strong>e of this chapter is as follows. In Section 5.1, we briefly discuss the various <strong>in</strong>gredientsenter<strong>in</strong>g <strong>in</strong>to the calculation of the A(e, e ′ p) matrix element. The follow<strong>in</strong>g sectionspresent the results of our A(e, e ′ p) numerical calculations. In Section 5.2, we look <strong>in</strong>to how thesecond-order eikonal correction affects an <strong>in</strong>clusive quantity like the nuclear transparency. The<strong>in</strong>duced normal polarization P n is the subject of <strong>in</strong>vestigation <strong>in</strong> Section 5.3. This observableserves as a rigid test for models deal<strong>in</strong>g with FSI mechanisms, s<strong>in</strong>ce it is exactly zero <strong>in</strong> the absenceof FSI. The left-right asymmetry A LT , another quantity that reflects the sensitivity to thedifferent <strong>in</strong>gredients that enter <strong>in</strong>to our model calculations, is studied <strong>in</strong> Section 5.4. F<strong>in</strong>ally,we end this chapter with a discussion of the effect of the second-order eikonal corrections onA(e, e ′ p) differential cross sections <strong>in</strong> Section 5.5.77


5.1. The A(e, e ′ p) Matrix Element 785.1 The A(e, e ′ p) Matrix ElementFor the description of the A(e, e ′ p) reaction, we adopt the impulse approximation (IA), wherea quasifree s<strong>in</strong>gle-nucleon knockout reaction mechanism is assumed, <strong>and</strong> the <strong>in</strong>dependentnucleonpicture. With<strong>in</strong> this approach, the basic quantity to be computed is the transitionmatrix element [221, 232]∫〈J µ 〉 = d⃗r Ψ (−)⃗ k,ms(⃗r) Ĵ µ (⃗r) e i⃗q·⃗r φ α1 (⃗r) . (5.1)Here, φ α1 <strong>and</strong> Ψ (−)⃗ k,msare the relativistic bound-state <strong>and</strong> scatter<strong>in</strong>g wave functions, with α 1 thestate where<strong>in</strong> the struck proton resided <strong>and</strong> ⃗ k <strong>and</strong> m s the momentum <strong>and</strong> sp<strong>in</strong> of the ejectedproton. The relativistic bound-state wave function is obta<strong>in</strong>ed <strong>in</strong> the Hartree approximationto the σ − ω model [38] with the W1 parametrization for the different field strengths [193]<strong>and</strong> is discussed <strong>in</strong> more detail <strong>in</strong> Appendix B. The scatter<strong>in</strong>g wave function Ψ (−)⃗appearsk,mswith <strong>in</strong>com<strong>in</strong>g boundary conditions <strong>and</strong> is related by time reversal to the st<strong>and</strong>ard scatter<strong>in</strong>gwave function Ψ (+)⃗as determ<strong>in</strong>ed <strong>in</strong> Sections 2.2 <strong>and</strong> 2.5. Furthermore, Ĵ µ is the relativistick,msone-body current operator model<strong>in</strong>g the coupl<strong>in</strong>g between the virtual photon <strong>and</strong> a nucleonembedded <strong>in</strong> the medium. Throughout this chapter, we use the Coulomb gauge <strong>and</strong> the CC2form of Ĵ µ [233]. For more details on the A(e, e ′ p) model, we refer to [221].5.2 <strong>Nuclear</strong> TransparencyOne way to quantify the overall effect of FSI <strong>in</strong> A(e, e ′ p) processes is via the nuclear transparency.The measurements [214–216, 218, 222, 223] are commonly performed under quasielasticconditions. We obta<strong>in</strong> the theoretical transparencies by adopt<strong>in</strong>g similar expressions <strong>and</strong>cuts as <strong>in</strong> the experiments. Hence, the nuclear transparency is extracted from the computedA(e, e ′ p) differential cross sections on the basis of the follow<strong>in</strong>g ratio [87]T =∑ ∫α ∆ 3 p md⃗p m S α (⃗p m , E m , ⃗ k)∫∆ 3 p md⃗p m SPWIA α (⃗p m, E m ) . (5.2)c A∑αHere, S α is the reduced cross section for knockout from the shell αS α (⃗p m , E m , ⃗ k) =d 5 σ αdΩ pdɛ ′ dΩ ɛ ′ (e, e′ p)Kσ ep, (5.3)where K is a k<strong>in</strong>ematical factor <strong>and</strong> σ ep is the off-shell electron-proton cross section, whichis evaluated with the CC1 prescription of de Forest [233]. SPWIA α is the reduced cross sectionwith<strong>in</strong> the plane-wave impulse approximation (PWIA) <strong>in</strong> the nonrelativistic limit. This limitis accomplished by nullify<strong>in</strong>g all sources of FSI mechanisms <strong>and</strong> neglect<strong>in</strong>g those contributions<strong>in</strong>troduced by the presence of negative-energy components <strong>in</strong> the relativistic bound-statewave functions [66]. Further, ∑ αextends over all occupied shells α <strong>in</strong> the target nucleus. Thephase-space volume <strong>in</strong> the miss<strong>in</strong>g momentum ∆ 3 p m is def<strong>in</strong>ed by the cut |p m | ≤ 300 MeV/c.


Chapter 5. Second-Order Eikonal Corrections <strong>in</strong> A(e, e ′ p) Observables 79F<strong>in</strong>ally, the A-dependent factor c A corrects <strong>in</strong> a phenomenological way for the effect of shortrangemechanisms. It accounts for the fact that short-range correlations move a fraction of thes<strong>in</strong>gle-particle strength to higher miss<strong>in</strong>g energies <strong>and</strong> momenta <strong>and</strong>, hence, beyond the rangescovered <strong>in</strong> ∑ ∫α ∆ 3 p md⃗p m of Eq. (5.2). We <strong>in</strong>troduce the c A <strong>in</strong> the denom<strong>in</strong>ator of Eq. (5.2) becausethe data have undergone a similar rescal<strong>in</strong>g. The adopted values for c A are 0.9 ( 12 C) <strong>and</strong>0.82 ( 56 Fe).Transparencies have been calculated for the nuclei 12 C <strong>and</strong> 56 Fe at planar <strong>and</strong> constant(⃗q, ω) k<strong>in</strong>ematics compatible with the phase space covered <strong>in</strong> the experiments. For the opticalpotential, the EDAD1 parametrization of Ref. [46] was used.In Fig. 5.1 the ROMEA <strong>and</strong> SOROMEA results are displayed as a function of the fourmomentumtransfer Q 2 <strong>and</strong> compared to the data. Not surpris<strong>in</strong>gly, at high Q 2 , the ROMEA<strong>and</strong> SOROMEA predictions practically co<strong>in</strong>cide <strong>and</strong> the second-order eikonal effects grow withdecreas<strong>in</strong>g Q 2 . At Q 2 = 1.7 (GeV/c) 2 , the ROMEA <strong>and</strong> SOROMEA transparencies agree to1%; while at Q 2 = 0.3 (GeV/c) 2 , the difference has risen to 3% for 56 Fe <strong>and</strong> 5% for 12 C. Theenhancement of the nuclear transparency due to the second-order eikonal corrections is ratherlimited, even for values of the four-momentum transfer as low as Q 2 = 0.2 (GeV/c) 2 . Boththe ROMEA <strong>and</strong> the SOROMEA predictions slightly underestimate the measurements. Thesecond-order corrections move the predictions somewhat closer to the Q 2 = 0.34 (GeV/c) 2data po<strong>in</strong>t.Further, Fig. 5.1 also displays the relativistic distorted-wave impulse approximation(RDWIA) results of Ref. [87]. The RDWIA framework was implemented by the Madrid-Sevillagroup [234] <strong>and</strong> is similar to our (SO)ROMEA approach <strong>in</strong> that both models compute the effectof the FSI with the aid of proton-nucleus optical potentials; but the RDWIA model relies on apartial-wave expansion of the exact scatter<strong>in</strong>g wave function. In order to make the comparisonwith the (SO)ROMEA results as mean<strong>in</strong>gful as possible, all the <strong>in</strong>gredients <strong>in</strong> the A(e, e ′ p)calculations not related to the FSI, such as the bound-state wave functions <strong>and</strong> the current operator,were kept identical. Moreover, the same EDAD1 parametrization of the optical potentialwas used. For high Q 2 , the RDWIA model predicts a few % more absorption than (SO)ROMEA,but the Q 2 dependence is quite similar down to Q 2 ≈ 0.7 (GeV/c) 2 . The low Q 2 dependence ofthe RDWIA transparencies is clearly steeper, even though the second-order corrections slightlyimprove the similarity <strong>in</strong> Q 2 dependence between the eikonal <strong>and</strong> the partial-wave models.The grow<strong>in</strong>g discrepancy between RDWIA <strong>and</strong> (SO)ROMEA with decreas<strong>in</strong>g Q 2 could be dueto the fact that the former does not adopt the EMA. On the whole, the SOROMEA transparencyresults are <strong>in</strong> slightly better agreement with the data than the RDWIA calculations, which systematicallyunderestimate the data by roughly 5–10%.In Fig. 5.2 the different model predictions of the attenuation for the <strong>in</strong>dividual shells <strong>in</strong> 12 C<strong>and</strong> 56 Fe are compared. These numbers are computed accord<strong>in</strong>g to the def<strong>in</strong>ition of Eq. (5.2)without perform<strong>in</strong>g the sum over the states α. The different s<strong>in</strong>gle-particle shells have verydifferent spatial characteristics. Consequently, by <strong>in</strong>vestigat<strong>in</strong>g the attenuation for each <strong>in</strong>di-


5.2. <strong>Nuclear</strong> Transparency 80112 C0.8Transparency0.60.40.656 Fe0.40.2SOROMEAROMEARDWIA0.2 0.3 0.4 0.5 0.6 0.7 0.80.9 1 2Q 2 (GeV 2 /c 2 )Figure 5.1 <strong>Nuclear</strong> transparencies versus Q 2 for A(e, e ′ p) reactions <strong>in</strong> quasielastic k<strong>in</strong>ematics. TheSOROMEA (dashed l<strong>in</strong>es) are compared to the ROMEA (solid l<strong>in</strong>es) results. The RDWIA calculations ofRef. [87] are also shown (dot-dashed curves). The EDAD1 potential [46] has been employed <strong>in</strong> both the(SO)ROMEA <strong>and</strong> RDWIA formalisms. Data po<strong>in</strong>ts are from Refs. [222] (open squares), [214, 215] (opentriangles), [216, 218] (solid triangles), <strong>and</strong> [223] (open diamonds).vidual shell <strong>in</strong> the target nucleus, the radial dependence of the FSI mechanisms can be studied.As expected, the calculations predict stronger attenuation for proton emission from a level witha larger fraction of its density <strong>in</strong> the nuclear <strong>in</strong>terior. Furthermore, the SOROMEA approachpredicts consistently less absorption than ROMEA for all shells. F<strong>in</strong>ally, it should be notedthat the discrepancy between RDWIA <strong>and</strong> (SO)ROMEA at low Q 2 is most evident for the 1s 1/2orbital. It is this same shell that almost entirely accounts for the strong rise of the 12 C RDWIAtransparency with decreas<strong>in</strong>g Q 2 . For the 56 Fe case, the RDWIA <strong>and</strong> (SO)ROMEA predictionsdiverge most strongly for the 1p 1/2 , 2s 1/2 , <strong>and</strong> 1s 1/2 shells.To put the second-order eikonal effects <strong>in</strong>to perspective, Fig. 5.3 shows the impact of thesp<strong>in</strong>-orbit V so ( ⃗ b, z) ⃗σ · ⃗b × ⃗ k <strong>and</strong> Darw<strong>in</strong> V so ( ⃗ b, z) (−ikz) terms on the nuclear transparency.Especially the Darw<strong>in</strong> term is important, with its effect be<strong>in</strong>g about three times larger than theimpact of the second-order eikonal effects. Even though the <strong>in</strong>clusion of the sp<strong>in</strong>-orbit term


Chapter 5. Second-Order Eikonal Corrections <strong>in</strong> A(e, e ′ p) Observables 8110.8SOROMEAROMEARDWIA0.501s 1/20.50.601p 3/2Transparency0.410.81s 1/2Transparency0.500.500.51p 1/21d 5/20.600.52s 1/20.41p 3/2Q 2 (GeV 2 /c 2 )00.51d 3/20.20.2 0.3 0.4 0.5 0.6 0.7 0.80.9 1 21f 7/2Q 2 (GeV 2 /c 2 )00.2 0.3 0.4 0.5 0.6 0.7 0.80.9 1 2Figure 5.2 The Q 2 dependence of the computed nuclear transparency for the s<strong>in</strong>gle-particle orbits asobta<strong>in</strong>ed <strong>in</strong> the SOROMEA, ROMEA, <strong>and</strong> RDWIA approaches with the EDAD1 optical potential ofRef. [46]. The left (right) panel shows the results for 12 C ( 56 Fe).partially undoes the reduction of the nuclear transparency due to the Darw<strong>in</strong> term, the globaleffect of sp<strong>in</strong>-orbit <strong>and</strong> Darw<strong>in</strong> term is still substantial. Further, Fig. 5.3 also displays results forthe EDAI parametrization of the optical potential. Clearly, both the predicted Q 2 dependence<strong>and</strong> the value of the transparency heavily depend on whether an A-dependent (EDAD1) orA-<strong>in</strong>dependent (EDAI) fit for the optical potentials is chosen.S<strong>in</strong>ce the nuclear transparency is an <strong>in</strong>tegrated quantity, it may hide some of the subtletiespresent <strong>in</strong> the treatment of the FSI. Next, we put our second-order eikonal corrections to amore str<strong>in</strong>gent test by focus<strong>in</strong>g on quantities that are probably more sensitive to the details ofthe calculations.5.3 Induced Normal PolarizationAn observable that is particularly well suited to study FSI effects is the <strong>in</strong>duced normal polarizationP n = d5 σ (σ n =↑) − d 5 σ (σ n =↓)d 5 σ (σ n =↑) + d 5 σ (σ n =↓) , (5.4)where σ n denotes the sp<strong>in</strong> orientation of the ejectile <strong>in</strong> the direction orthogonal to the reactionplane. Indeed, <strong>in</strong> the one-photon exchange approximation, P n vanishes <strong>in</strong> the absence of FSI.Fig. 5.4 shows the miss<strong>in</strong>g momentum dependence of the <strong>in</strong>duced normal polarization forthe k<strong>in</strong>ematics of Ref. [235], correspond<strong>in</strong>g with Q 2 ≈ 0.5 (GeV/c) 2 . The calculations employed


5.3. Induced Normal Polarization 82Transparency10.80.6Full, EDAD1noSO, EDAD1noD, EDAD1noSO-noD, EDAD1Full, EDAI0.40.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1Q 2 (GeV 2 /c 2 )Figure 5.3 The impact of the sp<strong>in</strong>-orbit <strong>and</strong> Darw<strong>in</strong> terms on the computed 12 C transparency <strong>and</strong> thesensitivity to the adopted choice for the parametrization of the optical potential. The solid curve is thefull SOROMEA calculation with the EDAD1 optical potential. The effect of turn<strong>in</strong>g off the sp<strong>in</strong>-orbit<strong>and</strong> Darw<strong>in</strong> terms is shown by the dashed <strong>and</strong> dot-dashed l<strong>in</strong>es, respectively, while the long-dottedcurve is obta<strong>in</strong>ed by switch<strong>in</strong>g off both terms. The short-dotted l<strong>in</strong>e displays the full SOROMEA resultus<strong>in</strong>g the EDAI optical potential. The data are from Refs. [214–216, 218, 222, 223].the energy-dependent A-<strong>in</strong>dependent EDAI potential of Ref. [46]. The ROMEA results are <strong>in</strong>l<strong>in</strong>e with the RDWIA calculations of the Madrid-Sevilla group [236] <strong>and</strong> the overall agreementwith the data is excellent. The second-order eikonal corrections are most pronounced for the1s 1/2 level. For miss<strong>in</strong>g momenta p m > 125 MeV/c, they reduce the magnitude of the P n forthe 1s 1/2 state by roughly 20%, thereby result<strong>in</strong>g <strong>in</strong> a marg<strong>in</strong>ally better agreement with thehighest p m data po<strong>in</strong>t. For 1p 3/2 knockout, on the other h<strong>and</strong>, the effect of the second-ordereikonal corrections does not exceed the 5% range.The <strong>in</strong>clusion of the second-order eikonal effects is particularly visible at high miss<strong>in</strong>g momentum,a region where other compet<strong>in</strong>g mechanisms also become more prom<strong>in</strong>ent. The qualitativebehavior of the meson-exchange <strong>and</strong> ∆-isobar currents, for <strong>in</strong>stance, is alike [237]. Atlow miss<strong>in</strong>g momenta p m ≤ 200 MeV/c, the <strong>in</strong>duced normal polarization P n is relatively <strong>in</strong>sensitiveto the two-body currents; whereas at higher miss<strong>in</strong>g momenta, sizable contributions fromthe meson-exchange <strong>and</strong> isobar currents are predicted. The <strong>in</strong>fluence of the meson <strong>and</strong> isobardegrees of freedom is also stronger for knockout from the 1s 1/2 shell than for 1p 3/2 knockout.In Fig. 5.4 also calculations without the sp<strong>in</strong>-orbit part V so ( ⃗ b, z) ⃗σ · ⃗b × ⃗ k are shown. Theyillustrate that the sp<strong>in</strong>-orbit distortion is the largest source of P n . Hence, a correct <strong>in</strong>clusionof this term is essential. Moreover, P n proves to be rather sensitive to the choice of opticalpotential [236].


Chapter 5. Second-Order Eikonal Corrections <strong>in</strong> A(e, e ′ p) Observables 830.50-0.5P nROMEASOROMEAROMEA-noSO 1pSOROMEA-noSO 3/20-0.512 C1s 1/2-10 50 100 150 200 250 300p m(MeV/c)Figure 5.4 Induced normal polarization P n for proton knockout from the 1p 3/2 (upper panel) <strong>and</strong>1s 1/2 (lower panel) shell <strong>in</strong> the 12 C(e, e ′ ⃗p) reaction. The k<strong>in</strong>ematics is determ<strong>in</strong>ed by beam energyɛ = 579 MeV, momentum transfer q = 760 MeV/c, energy transfer ω = 292 MeV, <strong>and</strong> azimuthal angleφ = 180 ◦ . The solid (dashed) curves represent ROMEA (SOROMEA) calculations. The dot-dashed(dotted) curves refer to predictions obta<strong>in</strong>ed with<strong>in</strong> the ROMEA (SOROMEA) frameworks, with thesp<strong>in</strong>-orbit term V so ( ⃗ b, z) ⃗σ ·⃗b × ⃗ k turned off. The data are from Ref. [235].5.4 Left-Right AsymmetryAnother A(e, e ′ p) observable which has been the subject of many <strong>in</strong>vestigations is the left-rightasymmetryA LT = d5 σ (φ = 0 ◦ ) − d 5 σ (φ = 180 ◦ )d 5 σ (φ = 0 ◦ ) + d 5 σ (φ = 180 ◦ ) . (5.5)The subscript LT <strong>in</strong>dicates that this quantity is closely related to the longitud<strong>in</strong>al-transverseresponse function. The experimental determ<strong>in</strong>ation of A LT is much less challeng<strong>in</strong>g than extract<strong>in</strong>gan absolute cross section or an effective response function. From a theoretical vantagepo<strong>in</strong>t, this ratio has the advantage of be<strong>in</strong>g <strong>in</strong>dependent of the spectroscopic factors. Furthermore,it is very well suited to scrut<strong>in</strong>ize different <strong>in</strong>gredients that enter various modelcalculations. Indeed, many subtleties that rema<strong>in</strong> concealed <strong>in</strong> other observables may becomeprom<strong>in</strong>ent <strong>in</strong> the A LT asymmetry.Fig. 5.5 presents our A LT predictions for the removal of 1p-shell protons <strong>in</strong> 16 O as a functionof the miss<strong>in</strong>g momentum <strong>in</strong> the k<strong>in</strong>ematics of Refs. [238,239]. The FSI shift the dip <strong>in</strong> A LT


5.5. Differential Cross Section 84which is located at p m ≈ 400 MeV/c <strong>in</strong> the relativistic PWIA (RPWIA), to lower values of themiss<strong>in</strong>g momentum. This shift is essential to describe the data at p m ≈ 350 MeV/c. The exactp m location <strong>and</strong> height of the ripple, however, are affected by many <strong>in</strong>gredients of the calculations,such as the current operator, bound-state wave function, <strong>and</strong> optical potential [239]. Ascan be <strong>in</strong>ferred from Fig. 5.5, the second-order eikonal corrections affect the height, but not thep m position of the ripple. For comparison reasons, we have chosen to only depict the resultsof our (SO)ROMEA calculations with<strong>in</strong> the so-called noSV approximation. In this approximation,the dynamical enhancement of the lower component of the scatter<strong>in</strong>g wave (2.43) due tothe V s − V v term is omitted. As such, our calculations employ the same set of basel<strong>in</strong>e optionsas the EMAf-noSV predictions by the Madrid-Sevilla group. The EMAf-noSV approachis an RDWIA calculation which adopts the EMA <strong>in</strong> comb<strong>in</strong>ation with the noSV approximation.The second-order effect clearly <strong>in</strong>creases the height of the oscillation <strong>in</strong> A LT <strong>and</strong> br<strong>in</strong>gsthe eikonal calculations <strong>in</strong> excellent agreement with the correspond<strong>in</strong>g partial-wave predictionEMAf-noSV. F<strong>in</strong>ally, the comparison with the full (SO)ROMEA <strong>and</strong> RDWIA calculationsdemonstrates that the <strong>in</strong>clusion of the dynamical enhancement is also an important aspect <strong>in</strong>the description of the A LT data.Next, we discuss some asymmetry results obta<strong>in</strong>ed for the 12 C nucleus by Dutta et al. [218].Their 12 C(e, e ′ p) experiment <strong>in</strong> quasiperpendicular k<strong>in</strong>ematics was performed with the aim ofextract<strong>in</strong>g nuclear transparencies, thereby measur<strong>in</strong>g reduced cross sections. The correspond<strong>in</strong>gleft-right asymmetry for reduced cross sections readsa LT = S (φ = 0◦ ) − S (φ = 180 ◦ )S (φ = 0 ◦ ) + S (φ = 180 ◦ ) , (5.6)with S def<strong>in</strong>ed by Eq. (5.3). Fig. 5.6 displays the data of Ref. [218] as well as our RPWIA,ROMEA, <strong>and</strong> SOROMEA calculations for the reduced asymmetry at three different Q 2 values.In accordance with the A LT predictions, the <strong>in</strong>clusion of FSI is crucial to reliably reproduce theexperimental asymmetry. Indeed, the FSI shift the p m ≈ 400 MeV/c dip <strong>in</strong> the RPWIA a LTto lower miss<strong>in</strong>g momenta. As was the case for the 16 O asymmetry of Fig. 5.5, the secondordereikonal corrections yield a larger dip <strong>in</strong> the asymmetry. The net effect of the eikonalcorrections, however, dim<strong>in</strong>ishes with ris<strong>in</strong>g Q 2 . The calculations are <strong>in</strong> l<strong>in</strong>e with the RDWIAresults of Ref. [240] <strong>and</strong> describe the 1p 3/2 data reasonably well. For the 1s 1/2 shell, on the otherh<strong>and</strong>, the ROMEA calculations fail to describe the p m distribution of the reduced asymmetry,<strong>in</strong> particular at the higher Q 2 values. Moreover, the second-order eikonal corrections seemto move the theoretical curves even further away from the 1s 1/2 data. The flatten<strong>in</strong>g of a LTas Q 2 <strong>in</strong>creases is not yet fully understood, but might <strong>in</strong>dicate greater contam<strong>in</strong>ation by themult<strong>in</strong>ucleon cont<strong>in</strong>uum.5.5 Differential Cross SectionIn Fig. 5.7 16 O(e, e ′ p) cross-section results are displayed for the k<strong>in</strong>ematics of Fig. 5.5. The spectroscopicfactors, which normalize the calculations to the data, were determ<strong>in</strong>ed by perform<strong>in</strong>g


Chapter 5. Second-Order Eikonal Corrections <strong>in</strong> A(e, e ′ p) Observables 850.50ROMEA-noSVSOROMEA-noSVEMAf-noSVRDWIARPWIA-0.5A LT-11p 1/20-0.51p 3/2-10 50 100 150 200 250 300 350 400 450p m(MeV/c)Figure 5.5 The left-right asymmetry A LT for the 16 O(e, e ′ p) experiment of [238]. The k<strong>in</strong>ematics wasɛ = 2.442 GeV, q = 1 GeV/c, <strong>and</strong> ω = 445 MeV (i.e., Q 2 = 0.8 (GeV/c) 2 ). The solid (dashed) l<strong>in</strong>es showthe results of the ROMEA-noSV (SOROMEA-noSV) calculations, which differ from the full calculations<strong>in</strong> that the dynamical enhancement of the lower component of the scatter<strong>in</strong>g wave function is neglected.The dot-dashed curves present the results from an RDWIA calculation where the sp<strong>in</strong>or distortions <strong>in</strong>the scattered wave are neglected, while the long-dotted curves display the full RDWIA predictions. Allcalculations use the EDAI version for the optical potentials [46]. The short-dotted curves represent theRPWIA results. The data po<strong>in</strong>ts are from Ref. [238].a χ 2 fit to the data <strong>and</strong> are summarized <strong>in</strong> Table 5.1. The RDWIA spectroscopic factors are 5–10% higher than the (SO)ROMEA ones. Both our (SO)ROMEA calculations <strong>and</strong> the RDWIApredictions of the Madrid-Sevilla group do a very good job of represent<strong>in</strong>g the data over theentire p m range. For miss<strong>in</strong>g momenta |p m | ≤ 250 MeV/c, the (SO)ROMEA calculations are<strong>in</strong> excellent agreement with the RDWIA ones. As for the other observables, the impact of thesecond-order eikonal corrections on the computed differential cross section is no more than afew percent for p m below the Fermi momentum, but grows to as much as 30% at higher miss<strong>in</strong>gmomenta. The <strong>in</strong>clusion of the second-order effects improves the agreement with the RDWIAcalculations at these high miss<strong>in</strong>g momenta. Results for the effective response functions R L ,R T , R T L , <strong>and</strong> R T T are not shown, but the effect of the second-order eikonal corrections is similarto the effect on the differential cross section.


5.5. Differential Cross Section 860.2Q 2 = 0.6 (GeV/c) 2 Q 2 = 1.3 (GeV/c) 2 Q 2 = 1.8 (GeV/c) 20-0.2a LT-0.4-0.60.21p 3/2RPWIAROMEASOROMEA0-0.2-0.4-0.61s 1/20 200 400200 400200 400p m(MeV/c)Figure 5.6 The left-right asymmetry for reduced cross sections for the 12 C(e, e ′ p) experiment of [218] atQ 2 = 0.6, 1.3, <strong>and</strong> 1.8 (GeV/c) 2 . The solid (dashed) curves refer to ROMEA (SOROMEA) calculations,us<strong>in</strong>g the EDAD1 optical potential [46]. The dot-dashed l<strong>in</strong>es show the RPWIA predictions.RPWIA ROMEA SOROMEA RDWIA1p 3/2 0.55 0.84 0.83 0.921p 1/2 0.47 0.75 0.74 0.78Table 5.1 The spectroscopic factors for the 16 O(e, e ′ p) reaction of Ref. [238], as obta<strong>in</strong>ed with a χ 2 procedure.Fig. 5.8 surveys the Q 2 dependence of the second-order eikonal effect on the 16 O(e, e ′ p)differential cross section. The ratio of the SOROMEA to the ROMEA calculations is shown as afunction of the miss<strong>in</strong>g momentum p m for two Q 2 values. The Q 2 = 0.8 (GeV/c) 2 k<strong>in</strong>ematics isas <strong>in</strong> Figs. 5.5 <strong>and</strong> 5.7, while the Q 2 = 0.3 (GeV/c) 2 k<strong>in</strong>ematics corresponds to the experimentby Ch<strong>in</strong>itz et al. [241]. For both Q 2 values, the differences between the SOROMEA <strong>and</strong> theROMEA predictions are more prom<strong>in</strong>ent at higher miss<strong>in</strong>g momenta. On the whole, Q 2 =0.3 (GeV/c) 2 leads to the strongest second-order effects; but even at this low Q 2 value, theeffect is at most 15% for p m below the Fermi momentum.


Chapter 5. Second-Order Eikonal Corrections <strong>in</strong> A(e, e ′ p) Observables 871d 5 σ/dε'dΩ ε'dΩ p(nb MeV -1 sr -2 )10 -110 -210 -3110 -1ROMEASOROMEARDWIARPWIA1p 3/210 -2p m(MeV/c)1p 1/210 -3-400 -300 -200 -100 0 100 200 300 400Figure 5.7 16 O(e, e ′ p) cross sections compared to ROMEA, SOROMEA, RDWIA, RPWIA calculations atthe quasiperpendicular k<strong>in</strong>ematics of Fig. 5.5. The calculations use the optical potential EDAI [46]. Thedata are from Ref. [238]. The follow<strong>in</strong>g convention is adopted: positive (negative) p m corresponds toφ = 180 ◦ (φ = 0 ◦ ).


5.5. Differential Cross Section 88d 5 σ(SOROMEA)/d 5 σ(ROMEA)1.21.110.9Q 2 = 0.3 (GeV/c) 2Q 2 = 0.8 (GeV/c) 21p 3/21p 1/20.8-250 -200 -150 -100 -50 0 50 100 150 200 250p m(MeV/c)Figure 5.8 The SOROMEA/ROMEA differential cross-section ratio for the 16 O(e, e ′ p) reaction <strong>in</strong>quasiperpendicular k<strong>in</strong>ematics at two values of Q 2 . The red (blue) curves refer to Q 2 =0.3 (0.8) (GeV/c) 2 .


Chapter 6Conclusions <strong>and</strong> OutlookIn this work, we have outl<strong>in</strong>ed a relativistic <strong>and</strong> cross-section factorized framework to describequasielastic A(p, pN) reactions. The model relies on the impulse approximation <strong>and</strong> the boundstatewave functions are obta<strong>in</strong>ed from a mean-field approximation to the σ − ω model [38].Relativity is accomodated <strong>in</strong> both the dynamics <strong>and</strong> the k<strong>in</strong>ematics. To model the propagationof the nucleons through the nuclear medium, a relativistic framework based on the eikonal approximation[54,55] has been developed. The eikonal approximation is a semiclassical method,which f<strong>in</strong>ds its orig<strong>in</strong>s <strong>in</strong> optics, <strong>and</strong> its applicability is restricted to small-angle scatter<strong>in</strong>g.This eikonal model provides a common framework to describe a variety of nuclear reactions<strong>and</strong> has been applied by our research group to A(e, e ′ p) [56, 57, 59, 87, 88], A(ν, ν ′ N), A(ν, ν ′ N),A(ν l , l − N), A(ν l , l + N) [242, 243], <strong>and</strong> A(γ, πN) [244] reactions. Here, however, the ma<strong>in</strong> focusis on A(p, pN) reactions. With three nucleons subject to attenuation effects, this reactionprovides an excellent test<strong>in</strong>g ground for the relativistic eikonal approximation.Our theoretical eikonal framework is very flexible as it can be used either <strong>in</strong> conjunctionwith relativistic optical potentials (ROMEA) or with<strong>in</strong> a Glauber multiple-scatter<strong>in</strong>g approach(RMSGA), which are two substantially different techniques to deal with the IFSI. Thanks tothe freedom of choice between these two techniques, our model is expected to be applicable atboth <strong>in</strong>termediate <strong>and</strong> high <strong>in</strong>cident energies.The ROMEA framework is similar to the DWIA approach, <strong>in</strong> which most theoretical workon A(p, pN) reactions was performed over the last number of decades, <strong>in</strong> the sense that bothof them <strong>in</strong>corporate IFSI <strong>in</strong> terms of optical potentials. These optical potentials are usuallyconstructed by global fits to nucleon-nucleus elastic scatter<strong>in</strong>g data. However, whereas theDWIA typically adopts a partial-wave expansion of the exact solution to the scatter<strong>in</strong>g problem,the ROMEA framework relies on the eikonal approximation to calculate the scatter<strong>in</strong>g wavefunction. The eikonal approach is particularly convenient at higher energies, where approachesrely<strong>in</strong>g on partial-wave expansions become <strong>in</strong>creas<strong>in</strong>gly impractical.For k<strong>in</strong>etic energies exceed<strong>in</strong>g 1 GeV, the elementary nucleon-nucleon scatter<strong>in</strong>g cross sectionsbecome <strong>in</strong>creas<strong>in</strong>gly <strong>in</strong>elastic <strong>and</strong> diffractive, <strong>and</strong> a treatment of the IFSI that relies on anoptical-potential approach can no longer be justified. In this high-energy region, the RMSGA89


90framework, which is a multiple-scatter<strong>in</strong>g extension of the eikonal approximation, offers avalid <strong>and</strong> economical alternative. In such a Glauber approach, the effects of IFSI are calculateddirectly from the nucleon-nucleon scatter<strong>in</strong>g data through the <strong>in</strong>troduction of a profilefunction. Thus, the major difference between both approaches is that the ROMEA frameworkdescribes the IFSI <strong>in</strong> terms of a nucleon-nucleus picture, whereas the RMSGA framework isessentially a nucleon-nucleon model. For k<strong>in</strong>etic energies above 1 GeV, the latter method isrecommended; while the former is more suitable at lower energies.In the calculations, the entire effect of the IFSI is accounted for by the IFSI factor. Theproperties of this function have been <strong>in</strong>vestigated for quasielastic proton scatter<strong>in</strong>g from 12 C,16 O, <strong>and</strong> 40 Ca at an <strong>in</strong>cident proton energy of 1 GeV, correspond<strong>in</strong>g with the k<strong>in</strong>ematics ofthe PNPI experiment of Ref. [114]. Not surpris<strong>in</strong>gly, the strongest attenuation occurs <strong>in</strong> thenuclear <strong>in</strong>terior, <strong>and</strong> heavier target nuclei are found to <strong>in</strong>duce larger IFSI effects. Also, thesurface-peaked character of the A(p, pN) reaction was clearly established to be a consequenceof the IFSI. In this experiment, as <strong>in</strong> many other A(p, pN) reactions, one faces the situation<strong>in</strong> which one of the ejectiles is relatively slow. Consequently, the Glauber multiple-scatter<strong>in</strong>gapproach is not applicable, even though the energies of the imp<strong>in</strong>g<strong>in</strong>g proton <strong>and</strong> the otherejectile are sufficiently large. The ROMEA calculations, on the other h<strong>and</strong>, give rise to realisticIFSI factors, with the different types of optical-potential sets conta<strong>in</strong>ed <strong>in</strong> Ref. [46] yield<strong>in</strong>gcomparable results.Next, the cross-section calculations for the PNPI experiment have been compared to thedata. The RMSGA approach fails to give an adequate description of the data because of the lowk<strong>in</strong>etic energy of the ejected nucleon, but the ROMEA calculations reproduce the shapes of themeasured differential cross sections reasonably well. Moreover, comparison with the RPWIApredictions illustrates the twofold effect of the IFSI. Besides the attenuation of the RPWIA crosssection, the IFSI also lead to a shift <strong>in</strong> the momentum distribution. The ROMEA model hasalso been used to calculate cross sections for the k<strong>in</strong>ematics of other experiments: 40 Ca(p, 2p)scatter<strong>in</strong>g at 460 MeV [115], 16 O(p, 2p) <strong>and</strong> (p, pn) scatter<strong>in</strong>g at 505 <strong>and</strong> 200 MeV [8, 116], <strong>and</strong>4 He(p, 2p) scatter<strong>in</strong>g at 250 MeV [9]. A fair description of the data is obta<strong>in</strong>ed for the differentk<strong>in</strong>ematics, thereby provid<strong>in</strong>g support for the wide applicability range of our model.The RMSGA approach, which was deemed unsuitable for the experiments discussed before,has been used to account for the IFSI <strong>in</strong> the study of A(p, 2p) nuclear transparenciesat high energies. To expla<strong>in</strong> the oscillatory energy dependence of the nuclear transparencydata [124–126], the Ralston-Pire model for the pp scatter<strong>in</strong>g amplitude [144] was implemented<strong>in</strong> our formalism. In their model, the free pp scatter<strong>in</strong>g amplitude consists of a small-sizedquark-count<strong>in</strong>g <strong>and</strong> a normal-sized L<strong>and</strong>shoff contribution. The <strong>in</strong>terference between thesetwo components results <strong>in</strong> oscillations <strong>in</strong> the free pp cross section. Inside the nuclear medium,the L<strong>and</strong>shoff component is absorbed via the regular Glauber mechanism, while the quarkcount<strong>in</strong>gcomponent escapes with relatively small attenuation due to color transparency. OurA(p, 2p) framework was adjusted to <strong>in</strong>corporate this so-called “nuclear filter<strong>in</strong>g” phenomenon.


Chapter 6. Conclusions <strong>and</strong> Outlook 91To describe the phenomenon of CT <strong>in</strong> our model, we considered two different models: thequantum diffusion model of Farrar et al. [148] <strong>and</strong> the hadronic expansion model of Jenn<strong>in</strong>gs<strong>and</strong> Miller [149].To our knowledge, the detailed RMSGA calculations of the A(p, 2p) nuclear transparencyare the first of their k<strong>in</strong>d. Unfortunately, the RMSGA procedure entails the numerical calculationof an <strong>in</strong>volv<strong>in</strong>g multi-dimensional <strong>in</strong>tegral which tracks the effect of all collisions of the<strong>in</strong>com<strong>in</strong>g <strong>and</strong> outgo<strong>in</strong>g protons with the rema<strong>in</strong><strong>in</strong>g nucleons <strong>in</strong> the target nucleus. Thereby,each of the target nucleons acts as a scatter<strong>in</strong>g center <strong>and</strong> is represented by its own relativisticwave function. To reduce the computational cost of the RMSGA calculations, some additionalapproximations were made, <strong>in</strong>clud<strong>in</strong>g replac<strong>in</strong>g the <strong>in</strong>dividual wave functions of the spectatornucleons by the average proton or neutron density of the residual nucleus. The nucleartransparency predictions with the full <strong>and</strong> the approximated RMSGA approach agree at thefew percent level. Thus, to determ<strong>in</strong>e <strong>in</strong>tegrated quantities such as the nuclear transparency, avalid alternative for the computationally <strong>in</strong>tensive RMSGA framework is available.Our calculations were brought <strong>in</strong> qualitative agreement with the A(p, 2p) transparency datathrough the <strong>in</strong>troduction of the nuclear filter<strong>in</strong>g concept. The major conclusion of our studyof the A(p, 2p) nuclear transparency is that CT is <strong>in</strong>dispensable to elevate the predicted transparenciesto the level observed by the experiments. The quantitative agreement with the data,however, can be much improved <strong>and</strong> several uncerta<strong>in</strong>ties still surround the research <strong>in</strong>toA(p, 2p) nuclear transparencies.F<strong>in</strong>ally, we have studied the validity of the eikonal approximation for the description ofthe A(e, e ′ p) reaction at lower energies. To this end, we have developed an extension of theROMEA model which accounts for second-order eikonal corrections, dubbed SOROMEA. TheA(e, e ′ p) nuclear transparency calculations confirm the expected energy dependence of thesecond-order eikonal corrections: the effect <strong>in</strong>creases with decreas<strong>in</strong>g Q 2 . However, even atQ 2 ≈ 0.2 (GeV/c) 2 , the effect of the second-order eikonal corrections on the A(e, e ′ p) nucleartransparency is rather modest. Further, we have paid attention to the miss<strong>in</strong>g momentum dependenceof several A(e, e ′ p) observables like the <strong>in</strong>duced normal polarization P n , the left-rightasymmetry A LT , <strong>and</strong> the differential cross section. At low miss<strong>in</strong>g momenta, the difference betweenthe ROMEA <strong>and</strong> SOROMEA predictions is observed to be negligible. In the high-p mregion, however, the second-order eikonal corrections are significant with effects up to the30% level. Thereby, the calculations are brought <strong>in</strong> closer agreement with the data <strong>and</strong>/or theRDWIA calculations.Overall, our numerical calculations show that the effect of the second-order eikonal correctionson A(e, e ′ p) observables is rather limited for Q 2 ≥ 0.2 (GeV/c) 2 . Vary<strong>in</strong>g the adoptedchoice of optical potential, current operator, bound-state wave functions, . . . leads to comparableeffects.To conclude, we would like to propose some possibilities for future work. To allow for the<strong>in</strong>clusion of sp<strong>in</strong> dependence <strong>in</strong> the description of the IFSI, the exist<strong>in</strong>g A(p, pN) model should


92be transformed <strong>in</strong>to an amplitude factorized formalism. Such a framework could then be usedto calculate A(p, pN) sp<strong>in</strong> observables. The second-order extension of the eikonal frameworkmight also prove useful <strong>in</strong> this context.The study of the A(p, 2p) nuclear transparency is another research area that is obviously stillopen for improvement. Here, a hybrid model that comb<strong>in</strong>es a partonic <strong>and</strong> a hadronic po<strong>in</strong>t ofview might result <strong>in</strong> a better description of the data.


Appendix ANotations <strong>and</strong> ConventionsA.1 AbbreviationsBNLc.m.CTEAEFIEMAFLFSFSIHHCIAIFSIIPMISIJMLNCSENFPLCPNPIpQCDQCQCDQEDQHD(R)DWIARMSGARMSGA ′ROMEA(R)PWIASOROMEASPVATSRCTRIUMFBrookhaven National Laboratorycenter of masscolor transparencyeikonal approximationEnrico Fermi Instituteeffective momentum approximationFarrar, Liu, Frankfurt, <strong>and</strong> Strikman modelf<strong>in</strong>al-state <strong>in</strong>teractionshadron helicity conservationimpulse approximation<strong>in</strong>itial- <strong>and</strong> f<strong>in</strong>al-state <strong>in</strong>teractions<strong>in</strong>dependent-particle model<strong>in</strong>itial-state <strong>in</strong>teractionsJenn<strong>in</strong>gs <strong>and</strong> Miller modelL<strong>and</strong>shoffnuclear color screen<strong>in</strong>g effectnuclear filter<strong>in</strong>gpo<strong>in</strong>t-like configurationPetersburg <strong>Nuclear</strong> <strong>Physics</strong> Instituteperturbative quantum chromodynamicsquark-count<strong>in</strong>gquantum chromodynamicsquantum electrodynamicsquantum hadrodynamics(relativistic) distorted-wave impulse approximationrelativistic multiple-scatter<strong>in</strong>g Glauber approximationapproximated RMSGArelativistic optical model eikonal approximation(relativistic) plane-wave impulse approximationsecond-order relativistic optical model eikonal approximationscalar, pseudoscalar, vector, axial vector, tensorshort-range correlationsTRI-University Meson Facility93


A.2. Pauli Matrices 94A.2 Pauli MatricesThe sp<strong>in</strong> operator is def<strong>in</strong>ed as⃗σ = (σ x , σ y , σ z ) = σ i ⃗e i ,(A.1)where the Pauli matrices are given byσ x =( ) 0 11 0, σ y =( ) 0 −ii 0, σ z =( ) 1 00 −1The isosp<strong>in</strong> operator ⃗τ = (τ 1 , τ 2 , τ 3 ) has the same matrix representation.. (A.2)A.3 Dirac MatricesThe Dirac or γ matrices are def<strong>in</strong>ed by the anticommutation relations{γ µ , γ ν } = 2 g µν , (A.3)<strong>and</strong>, <strong>in</strong> the Dirac-Pauli representation, are given byγ 0 =( ) 1 00 −1, γ i =( 0)σi−σ i 0Other commonly used matrices are β <strong>and</strong> ⃗α, def<strong>in</strong>ed as. (A.4)β = γ 0 , ⃗α = γ 0 ⃗γ . (A.5)


Appendix BRelativistic Bound-State Wave FunctionsTraditionally, nuclear physics calculations relied on the nonrelativistic Schröd<strong>in</strong>ger equation. Innonrelativistic many-body theory, nuclei are described as bound states of nucleons <strong>in</strong>teract<strong>in</strong>gvia two- <strong>and</strong> three-body potentials. Although this approach has enjoyed many successes <strong>in</strong>describ<strong>in</strong>g the properties of nuclear matter, a relativistic treatment is preferable for severalreasons. First, the mesonic degrees of freedom can be <strong>in</strong>troduced at the start of the developmentof the model. Second, the pr<strong>in</strong>ciples of causality, retardation, <strong>and</strong> relativistic k<strong>in</strong>ematics can beeasily <strong>in</strong>corporated <strong>in</strong> a relativistic framework. Furthermore, the sp<strong>in</strong>-orbit <strong>in</strong>teraction needs tobe <strong>in</strong>serted by h<strong>and</strong> <strong>in</strong> nonrelativistic theories, but emerges automatically <strong>in</strong> relativistic models.Relativistic quantum field theories based on hadrons, known as quantum hadrodynamics(QHD), have been found to be very successful <strong>in</strong> describ<strong>in</strong>g the bulk <strong>and</strong> s<strong>in</strong>gle-particle propertiesof nuclear matter <strong>and</strong> f<strong>in</strong>ite nuclei [38,245–248]. The approach was orig<strong>in</strong>ally <strong>in</strong>troducedby Walecka roughly 30 years ago [36] <strong>and</strong> uses the nucleon (as a Dirac particle), the isoscalarscalarσ, <strong>and</strong> the isoscalar-vector ω mesons as the relevant degrees of freedom. In this so-calledσ − ω model, the Lagrangian density reads [37, 38]L 0 = ¯ψ (i̸ ∂ − M N ) ψ + 1 2 (∂ µφ ∂ µ φ − m 2 s φ 2 ) − 1 4 G µν G µν+ 1 2 m2 v V µ V µ − g v ¯ψγµ ψV µ + g s ¯ψψφ . (B.1)The nucleons (ψ) <strong>in</strong>teract with scalar fields (φ) through a Yukawa coupl<strong>in</strong>g ψψφ <strong>and</strong> with neutralvector fields (V µ ) that couple to the conserved baryon current ψγ µ ψ. Here, M N , m s , <strong>and</strong>m v are the nucleon, scalar meson, <strong>and</strong> vector meson masses, respectively, g s <strong>and</strong> g v are thescalar <strong>and</strong> vector coupl<strong>in</strong>gs to the nucleon, <strong>and</strong> G µν ≡ ∂ µ V ν − ∂ ν V µ is the vector meson fieldstrength. The scalar (φ) <strong>and</strong> vector (V µ ) fields are associated with the σ <strong>and</strong> ω mesons. Solv<strong>in</strong>gthe complete quantum field theory is far from trivial <strong>and</strong> to tackle the problem, the meson fieldoperators are usually replaced by their classical expectation values or “mean fields”. This constitutesthe so-called Hartree approximation [36]. In <strong>in</strong>f<strong>in</strong>ite matter, this amounts to 〈φ〉 ≡ φ 0<strong>and</strong> 〈V µ 〉 ≡ δ µ0 V 0 . In a Hartree-Fock approach, the result<strong>in</strong>g mean-field problem can then besolved iteratively. The model can be extended to also <strong>in</strong>clude the isovector-vector ρ <strong>and</strong> the95


96pseudo-scalar π mesons, as well as the coupl<strong>in</strong>g to the photon field [249].In an attempt to <strong>in</strong>clude the pr<strong>in</strong>ciples of quantum chromodynamics (QCD), various QHDmodels have been developed. In particular, the constriant of chiral symmetry plays an importantrole <strong>in</strong> hadronic physics at low energies <strong>and</strong>, consequently, there is a long history ofattempts to unite relativistic mean-field phenomenology based on hadrons with manifest chiralsymmetry. An overview of the successes <strong>and</strong> failures of the different models is given <strong>in</strong>Refs. [247,248]. Whereas <strong>in</strong> the orig<strong>in</strong>al Walecka model (B.1) no attempt was made to reconcilethe model with the spontaneously broken, approximate SU(2) L × SU(2) R chiral symmetry ofhadronic <strong>in</strong>teractions; <strong>in</strong> the model developed by Furnstahl et al. [193, 250, 251], a nonl<strong>in</strong>earrealization of chiral symmetry is adopted. This chiral effective field theory <strong>in</strong>cludes all the underly<strong>in</strong>gsymmetries of QCD, such as Lorentz covariance, parity conservation, time-reversal<strong>and</strong> charge-conjugation <strong>in</strong>variance, isosp<strong>in</strong> symmetry, unitarity, <strong>and</strong> gauge <strong>in</strong>variance. Moreover,it is based on a density-functional approach <strong>in</strong> which higher-order many-body correctionsare treated approximately, which is equivalent to a Hartree calculation [251]. The results arecalculated at one-baryon-loop order, which is corresponds with the Dirac-Hartree approximation[37].The Furnstahl model leads to a static Dirac equation [38]Ĥ φ α (⃗r, ⃗σ) = E α φ α (⃗r, ⃗σ) ,(B.2)with eigenvalues E α <strong>and</strong> eigenfunctions φ α (⃗r, ⃗σ), <strong>and</strong> where the s<strong>in</strong>gle-particle Dirac Hamiltonianis given by [193]Ĥ = −i ⃗α · ⃗∇ + g v V 0 (⃗r) + 1 2 τ 3 g ρ b 0 (⃗r) + β (M N − g s Φ 0 (⃗r))+ 1 2 (1 + τ 3) e A 0 (⃗r) − i2M Nβ ⃗α · (fρ12 τ 3 g ρ⃗ ∇b0 (⃗r) + f v g v⃗ ∇V0 (⃗r) )+ 12MN2 (β s + β v τ 3 ) e ∇ 2 A 0 (⃗r) − i λ β e ⃗α ·2M ⃗∇A 0 (⃗r) .N(B.3)Here, Φ 0 , V 0 , b 0 , <strong>and</strong> A 0 refer to the σ, ω, ρ, <strong>and</strong> electromagnetic mean fields, respectively;<strong>and</strong> g s , g v , g ρ , <strong>and</strong> e are the correspond<strong>in</strong>g scalar, vector, rho, <strong>and</strong> electromagnetic coupl<strong>in</strong>gs.The coupl<strong>in</strong>g between the ρ (ω) meson <strong>and</strong> the nucleon is denoted by f ρ (f v ), while β s (β v )is the coupl<strong>in</strong>g for higher-order σN <strong>and</strong> σσ (ωN <strong>and</strong> ωω) <strong>in</strong>teractions. Furthermore, τ 3 is thethird component of the isosp<strong>in</strong> operator <strong>and</strong> λ ≡ 1 2 λ p (1 + τ 3 ) + 1 2 λ n (1 − τ 3 ) is the anomalousmagnetic moment, with λ p = 1.793 <strong>and</strong> λ n = −1.913. The pion field does not enter <strong>in</strong> theHartree approximation, if one assumes that the nuclear ground state is spherically symmetric<strong>and</strong> a parity eigenstate [252].For spherically symmetric potentials, the solutions φ α (⃗r, ⃗σ) to a s<strong>in</strong>gle-particle Dirac equationhave the form [253]φ α (⃗r, ⃗σ) ≡ φ nκm (⃗r, ⃗σ) =[i Gnκ(r)r− Fnκ(r)r]Y κm (Ω, ⃗σ), (B.4)Y −κm (Ω, ⃗σ)


Appendix B. Relativistic Bound-State Wave Functions 97where n denotes the pr<strong>in</strong>cipal <strong>and</strong> κ <strong>and</strong> m the generalized angular momentum quantum numbers.The Y ±κm are the sp<strong>in</strong> spherical harmonics <strong>and</strong> determ<strong>in</strong>e the angular <strong>and</strong> sp<strong>in</strong> parts ofthe wave function,Y κm (Ω, ⃗σ) = ∑ 〈〉1lm l2 m s | jmm l m swithY −κm (Ω, ⃗σ) = ∑ m l m s〈¯lml12 m s | jm〉Y lml (Ω) χ 12 ms(⃗σ) ,Y¯lml (Ω) χ 1ms(⃗σ) , (B.5)2j = |κ| − 1 2 , l = { κ, κ > 0−(κ + 1), κ < 0, ¯l = 2j − l ={ κ − 1, κ > 0−κ, κ < 0. (B.6)The Dirac equation (B.2) (or the Lagrangian of Eq. (B.1) solved <strong>in</strong> the relativistic Hartree approximation)leads to a set of coupled equations for the different fields Φ 0 , V 0 , b 0 , A 0 , G nκ , <strong>and</strong>F nκ . This set of equations is solved iteratively, start<strong>in</strong>g from an educated guess of the mesonpotentials. The solutions depend on the nucleon <strong>and</strong> meson masses <strong>and</strong> coupl<strong>in</strong>g constants. Inthis work, all bound-state wave functions are computed us<strong>in</strong>g the W1 parametrization set ofFurnstahl et al. [193].In Fig. B.1, the charge density <strong>and</strong> form factor of 40 Ca are shown. The results obta<strong>in</strong>edwith the W1 parametrization of Ref. [193] are compared with the predictions of Ref. [37]. Thetheoretical results agree reasonably well with the data of Ref. [254], <strong>and</strong> the correspondencebetween the theoretical <strong>and</strong> experimental average b<strong>in</strong>d<strong>in</strong>g energy per nucleon is also quitegood. Even though an IPM description of the four-nucleon system appears rather questionable,the bound-state wave functions of the 4 He nucleus are also calculated <strong>in</strong> the Furnstahlmodel [193]. After all, at present, realistic relativistic 4 He wave functions are not available,<strong>and</strong> a mean-field approach is the only viable option. The W1 parametrization reproduces thelow-momentum-transfer part of the 4 He charge form factor fairly well [221]. The discrepancybetween the calculation <strong>and</strong> the data at high momentum transfer is caused by the two-bodycharge operator contributions [255].The Fourier transform of the relativistic bound-nucleon wave function (B.4) is given by∫φ α (⃗p) =d⃗r e −i⃗p·⃗r φ α (⃗r) = (−i) l (2π) 3/2 [g nκ (p) Y κm (Ω p )−S κ f nκ (p) Y −κm (Ω p )], (B.7)with S κ = κ/ |κ|. The radial functions g nκ <strong>and</strong> f nκ <strong>in</strong> momentum space are obta<strong>in</strong>ed from theircounterparts <strong>in</strong> coord<strong>in</strong>ate space:g nκ (p) = if nκ (p) = i√2π√2π∫ ∞0∫ ∞0r 2 dr G nκ(r)rr 2 dr F nκ(r)rj l (pr) ,j¯l(pr) ,with j l (pr) the spherical Bessel function of the first k<strong>in</strong>d.(B.8a)(B.8b)


980.11ρ chg(e/fm 3 )0.080.060.0440 CaHorowitzW1|F(q)|10 -110 -20.0200 1 2 3 4 5r (fm)10 -30 0.5 1 1.5 2 2.5q (fm -1 )Figure B.1 The left (right) panel shows the charge density (form factor) of 40 Ca. The solid curve displaysthe prediction of Ref. [37]. The dashed curve is obta<strong>in</strong>ed with the W1 parametrization of Ref. [193].


Appendix CRadial <strong>and</strong> Polar-Angle Contribution tothe A(p, pN) Cross SectionThe differential A(p, pN) cross section (2.40) is proportional to the distorted momentum distributionρ D of Eq. (2.38). When approximat<strong>in</strong>g the completeness relation (2.17) as∑u(⃗p m , m s ) ū(⃗p m , m s ) ≈ 1 ,si.e., neglect<strong>in</strong>g the negative-energy term as <strong>in</strong> Section 2.1.2, this amounts to(d 5 ) Dσ∝ ∑ φdE k1 dΩ 1 dΩ D α 1φ D α 1.2m(C.1)(C.2)Thus, with D r (r) def<strong>in</strong>ed as∫D r (r) ≡ dΩ r 2 e −i⃗pm·⃗r φ α1 (⃗r) ŜIFSI (⃗r) ,(C.3)the functionδ r (r 1 ) ≡ ∑ [ ∫1 ∞dr D r (r)∆Rm 0( ∫ r 1∫ ∞dr D r (r) += ∑ m0r 1 +∆R∫ ∞0dr D r (r) −) ( ∫ r 1dr D r (r)0∫ ∞dr D r (r) +r 1 +∆R) ]dr D r (r)(Dr (r 1 ) φ D α 1+ φ D α 1D r (r 1 ) ) (C.4)represents the contribution of an <strong>in</strong>f<strong>in</strong>itesimal <strong>in</strong>terval <strong>in</strong> r around r 1 (<strong>in</strong>tegrated over the entire(θ, φ) range) to the A(p, pN) cross section. This procedure also enables us to estimate theaverage density seen through this reaction as∫ ∞0ρ (r) δ r (r) dr〈ρ〉 ≡ ∫ ∞. (C.5)0δ r (r) drIn a similar fashion, the functionδ r,θ (r 1 , θ 1 ) ≡ ∑ m(Dr,θ (r 1 , θ 1 ) φ D α 1+ φ D α 1D r,θ (r 1 , θ 1 ) ) , (C.6)99


100with∫D r,θ (r, θ) ≡dφ s<strong>in</strong> θ r 2 e −i⃗pm·⃗r φ α1 (⃗r) ŜIFSI (⃗r) ,(C.7)is the cross-section contribution of an <strong>in</strong>f<strong>in</strong>itesimal <strong>in</strong>terval <strong>in</strong> r <strong>and</strong> θ around r 1 <strong>and</strong> θ 1 .


Bibliography[1] G. Jacob <strong>and</strong> T. A. J. Maris, “Quasi-Free Scatter<strong>in</strong>g <strong>and</strong> <strong>Nuclear</strong> Structure,” Rev. Mod.Phys. 38, 121 (1966).[2] G. Jacob <strong>and</strong> T. A. J. Maris, “Quasi-Free Scatter<strong>in</strong>g <strong>and</strong> <strong>Nuclear</strong> Structure. II,” Rev. Mod.Phys. 45, 6 (1973).[3] P. Kitch<strong>in</strong>g, W. J. McDonald, T. A. J. Maris, <strong>and</strong> C. A. Z. Vasconcellos, “Recent Developments<strong>in</strong> Quasi-Free Nucleon-Nucleon Scatter<strong>in</strong>g,” Adv. Nucl. Phys. 15, 43 (1985).[4] N. S. Chant <strong>and</strong> P. G. Roos, “Distorted-wave impulse-approximation calculations forquasifree cluster knockout reactions,” Phys. Rev. C 15, 57 (1977).[5] S. Kull<strong>and</strong>er, F. Lemeilleur, P. U. Renberg, G. L<strong>and</strong>aud, J. Yonnet, B. Fagerström, A. Johansson,<strong>and</strong> G. Tibell, “(p, 2p) reactions at 600 MeV <strong>and</strong> the distorted-wave impulseapproximation,” Nucl. Phys. A 173, 357 (1971).[6] R. K. Bhowmik, C. C. Chang, P. G. Roos, <strong>and</strong> H. D. Holmgren, “The 7 Li(p, 2p) 5 He reactionat 100 MeV,” Nucl. Phys. A 226, 365 (1974).[7] R. Bengtsson, T. Berggren, <strong>and</strong> C. Gustafsson, “DWIA analysis of quasifree proton-protonscatter<strong>in</strong>g <strong>in</strong> nuclei at <strong>in</strong>termediate energies,” Phys. Rep. 41, 191 (1978).[8] P. Kitch<strong>in</strong>g, C. A. Miller, W. C. Olsen, D. A. Hutcheon, W. J. McDonald, <strong>and</strong> A. W. Stetz,“Quasi-free scatter<strong>in</strong>g of polarized protons,” Nucl. Phys. A 340, 423 (1980).[9] W. T. H. van Oers et al., “ 4 He(p, 2p) 3 H reaction at <strong>in</strong>termediate energies,” Phys. Rev. C25, 390 (1982).[10] D. S. Carman et al., “Quasifree <strong>in</strong>clusive <strong>and</strong> exclusive cross sections at 200 MeV,” Phys.Lett. B 452, 8 (1999).[11] N. S. Chant <strong>and</strong> P. G. Roos, “Sp<strong>in</strong> orbit effects <strong>in</strong> quasifree knockout reactions,” Phys.Rev. C 27, 1060 (1983).101


Bibliography 102[12] R. A. Arndt, I. I. Strakovsky, <strong>and</strong> R. L. Workman, “Nucleon-nucleon elastic scatter<strong>in</strong>g to3 GeV,” Phys. Rev. C 62, 034005 (2000).[13] Scatter<strong>in</strong>g Analysis Interactive Dial-<strong>in</strong> Program (SAID), http://gwdac.phys.gwu.edu/ .[14] J. A. McNeil, L. Ray, <strong>and</strong> S. J. Wallace, “Impulse approximation NN amplitudes forproton-nucleus <strong>in</strong>teractions,” Phys. Rev. C 27, 2123 (1983).[15] J. A. McNeil, J. R. Shepard, <strong>and</strong> S. J. Wallace, “Impulse-Approximation Dirac OpticalPotential,” Phys. Rev. Lett. 50, 1439 (1983).[16] J. R. Shepard, J. A. McNeil, <strong>and</strong> S. J. Wallace, “Relativistic Impulse Approximation forp-Nucleus Elastic Scatter<strong>in</strong>g,” Phys. Rev. Lett. 50, 1443 (1983).[17] C. J. Horowitz, “Relativistic Love-Franey model: Covariant representation of the NN<strong>in</strong>teraction for N-nucleus scatter<strong>in</strong>g,” Phys. Rev. C 31, 1340 (1985).[18] C. J. Horowitz <strong>and</strong> M. J. Iqbal, “Relativistic effects on sp<strong>in</strong> observables <strong>in</strong> quasielasticproton scatter<strong>in</strong>g,” Phys. Rev. C 33, 2059 (1986).[19] D. P. Murdock <strong>and</strong> C. J. Horowitz, “Microscopic relativistic description of proton-nucleusscatter<strong>in</strong>g,” Phys. Rev. C 35, 1442 (1987).[20] C. J. Horowitz <strong>and</strong> D. P. Murdock, “Quasielastic proton-nucleus scatter<strong>in</strong>g <strong>in</strong> a relativisticplane-wave impulse approximation,” Phys. Rev. C 37, 2032 (1988).[21] O. V. Maxwell, “Energy-dependent Lorentz covariant representation of the NN <strong>in</strong>teraction,”Nucl. Phys. A 600, 509 (1996).[22] O. V. Maxwell, “Lorentz covariant representation of the NN <strong>in</strong>teraction at energies above500 MeV,” Nucl. Phys. A 638, 747 (1998).[23] O. V. Maxwell <strong>and</strong> E. D. Cooper, “Elastic unitarity <strong>and</strong> the K-matrix <strong>in</strong> a Lorentz covariantrepresentation of the NN <strong>in</strong>teraction,” Nucl. Phys. A 656, 231 (1999).[24] E. D. Cooper <strong>and</strong> O. V. Maxwell, “The (p, 2p) reaction <strong>in</strong> the Dirac impulse approximation,”Nucl. Phys. A 493, 468 (1989).[25] O. V. Maxwell <strong>and</strong> E. D. Cooper, “Proton-<strong>in</strong>duced nucleon knockout from 40 Ca <strong>in</strong> theDirac impulse approximation,” Nucl. Phys. A 513, 584 (1990).[26] O. V. Maxwell <strong>and</strong> E. D. Cooper, “Proton knockout from 16 O <strong>and</strong> relativistic recoil effects<strong>in</strong> the Dirac impulse approximation,” Nucl. Phys. A 565, 740 (1993).[27] O. V. Maxwell <strong>and</strong> E. D. Cooper, “Proton knockout from 16 O at 504 MeV <strong>and</strong> limitationsof the DWIA,” Nucl. Phys. A 574, 819 (1994).


Bibliography 103[28] O. V. Maxwell <strong>and</strong> E. D. Cooper, “Interaction dependence <strong>in</strong> a relativistic DWIA descriptionof proton knockout,” Nucl. Phys. A 603, 441 (1996).[29] Y. Ikebata, “(p, 2p) reaction <strong>in</strong> zero-range <strong>and</strong> f<strong>in</strong>ite-range relativistic distorted-wave impulseapproximation,” Phys. Rev. C 52, 890 (1995).[30] J. Mano <strong>and</strong> Y. Kudo, “A Self-Consistent Description of (p, 2p) Reactions <strong>in</strong> a RelativisticF<strong>in</strong>ite-Range DWIA Model,” Prog. Theor. Phys. 100, 91 (1998).[31] C. A. Miller et al., “Sp<strong>in</strong>-dependent scatter<strong>in</strong>g of deeply bound nucleons,” Phys. Rev. C57, 1756 (1998).[32] R. Nevel<strong>in</strong>g, A. A. Cowley, G. F. Steyn, S. V. Förtsch, G. C. Hillhouse, J. Mano, <strong>and</strong> S. M.Wyngaardt, “Analyz<strong>in</strong>g power <strong>and</strong> cross section distributions of the knockout reaction208 Pb(⃗p, 2p) 207 Tl at an <strong>in</strong>cident energy of 202 MeV,” Phys. Rev. C 66, 034602 (2002).[33] D. L. Adams <strong>and</strong> M. Bleszynski, “On the relevance of the Dirac equation to the scatter<strong>in</strong>gof medium-energy nucleons from nuclei,” Phys. Lett. B 136, 10 (1984).[34] J. A. Tjon <strong>and</strong> S. J. Wallace, “General Lorentz-<strong>in</strong>variant representation of NN scatter<strong>in</strong>gamplitudes,” Phys. Rev. C 32, 1667 (1985).[35] B. I. S. van der Ventel <strong>and</strong> G. C. Hillhouse, “Sensitivity of exclusive proton knockout sp<strong>in</strong>observables to different Lorentz <strong>in</strong>variant representations of the NN <strong>in</strong>teraction,” Phys.Rev. C 69, 024618 (2004).[36] J. D. Walecka, “A theory of highly condensed matter,” Ann. Phys. (N.Y.) 83, 491 (1974).[37] C. J. Horowitz <strong>and</strong> B. D. Serot, “Self-consistent hartree description of f<strong>in</strong>ite nuclei <strong>in</strong> arelativistic quantum field theory,” Nucl. Phys. A 368, 503 (1981).[38] B. D. Serot <strong>and</strong> J. D. Walecka, “The Relativistic <strong>Nuclear</strong> Many Body Problem,” Adv. Nucl.Phys. 16, 1 (1986).[39] D. F. Jackson <strong>and</strong> T. Berggren, “Partial wave analysis of the (p, 2p) reaction,” Nucl. Phys.62, 353 (1965).[40] A. Nadasen, P. Schw<strong>and</strong>t, P. P. S<strong>in</strong>gh, W. W. Jacobs, A. D. Bacher, P. T. Debevec, M. D.Kaitchuck, <strong>and</strong> J. T. Meek, “Elastic scatter<strong>in</strong>g of 80-180 MeV protons <strong>and</strong> the protonnucleusoptical potential,” Phys. Rev. C 23, 1023 (1981).[41] P. Schw<strong>and</strong>t, H. O. Meyer, W. W. Jacobs, A. D. Bacher, S. E. Vigdor, M. D. Kaitchuck, <strong>and</strong>T. R. Donoghue, “Analyz<strong>in</strong>g power of proton-nucleus elastic scatter<strong>in</strong>g between 80 <strong>and</strong>180 MeV,” Phys. Rev. C 26, 55 (1982).[42] D. G. Madl<strong>and</strong>, Los Alamos National Laboratory Report LA-UR-87-3382, 1987 (unpublished).


Bibliography 104[43] E. D. Cooper, B. C. Clark, R. Kozack, S. Shim, S. Hama, J. I. Johansson, H. S. Sherif, R. L.Mercer, <strong>and</strong> B. D. Serot, “Global optical potentials for elastic p+ 40 Ca scatter<strong>in</strong>g us<strong>in</strong>g theDirac equation,” Phys. Rev. C 36, 2170 (1987).[44] E. D. Cooper, B. C. Clark, S. Hama, <strong>and</strong> R. L. Mercer, “Dirac-global fits to calcium elasticscatter<strong>in</strong>g data <strong>in</strong> the range 21-200 MeV,” Phys. Lett. B 206, 588 (1988).[45] S. Hama, B. C. Clark, E. D. Cooper, H. S. Sherif, <strong>and</strong> R. L. Mercer, “Global Dirac opticalpotentials for elastic proton scatter<strong>in</strong>g from heavy nuclei,” Phys. Rev. C 41, 2737 (1990).[46] E. D. Cooper, S. Hama, B. C. Clark, <strong>and</strong> R. L. Mercer, “Global Dirac phenomenology forproton-nucleus elastic scatter<strong>in</strong>g,” Phys. Rev. C 47, 297 (1993).[47] L. Antonuk, P. Kitch<strong>in</strong>g, C. A. Miller, D. A. Hutcheon, W. J. McDonald, G. C. Neilson,<strong>and</strong> W. C. Olsen, “An <strong>in</strong>vestigation of the 40 Ca(p, 2p) reaction with polarised protons,”Nucl. Phys. A 370, 389 (1981).[48] C. Samanta, N. S. Chant, P. G. Roos, A. Nadasen, J. Wesick, <strong>and</strong> A. A. Cowley, “Tests ofthe factorized distorted wave impulse approximation for (p, 2p) reactions,” Phys. Rev. C34, 1610 (1986).[49] A. A. Cowley, J. J. Lawrie, G. C. Hillhouse, D. M. Whittal, S. V. Förtsch, J. V. Pilcher,F. D. Smit, <strong>and</strong> P. G. Roos, “Quasifree knockout <strong>in</strong> 16 O(p, 2p) 15 N at an <strong>in</strong>cident energy of151 MeV,” Phys. Rev. C 44, 329 (1991).[50] A. A. Cowley, G. J. Arendse, J. A. St<strong>and</strong>er, <strong>and</strong> W. A. Richter, “Distortion <strong>in</strong> protonknockoutreactions,” Phys. Lett. B 359, 300 (1995).[51] A. A. Cowley et al., “Rescatter<strong>in</strong>g <strong>in</strong> knockout reactions as manifested <strong>in</strong> 40 Ca(p, p ′ p ′′ ) atan <strong>in</strong>cident energy of 392 MeV,” Phys. Rev. C 57, 3185 (1998).[52] G. C. Hillhouse, J. Mano, A. A. Cowley, <strong>and</strong> R. Nevel<strong>in</strong>g, “Relativistic predictions ofexclusive 208 Pb(⃗p, 2p) 207 Tl analyz<strong>in</strong>g powers at an <strong>in</strong>cident energy of 202 MeV,” Phys.Rev. C 67, 064604 (2003).[53] G. C. Hillhouse, J. Mano, S. M. Wyngaardt, B. I. S. van der Ventel, T. Noro, <strong>and</strong> K.Hatanaka, “Relativistic predictions of sp<strong>in</strong> observables for exclusive proton knockoutreactions,” Phys. Rev. C 68, 034608 (2003).[54] G. P. McCauley <strong>and</strong> G. E. Brown, “Inelastic Scatter<strong>in</strong>g of High Energy Nucleons by ComplexNuclei I: Semi-Classical Formalism,” Proc. Phys. Soc. London 71, 893 (1958).[55] C. J. Joacha<strong>in</strong>, Quantum Collision Theory (Elsevier, Amsterdam, 1975).[56] D. Debruyne, J. Ryckebusch, W. Van Nespen, <strong>and</strong> S. Janssen, “Relativistic eikonal approximation<strong>in</strong> high-energy A(e, e ′ p) reactions,” Phys. Rev. C 62, 024611 (2000).


Bibliography 105[57] D. Debruyne, J. Ryckebusch, S. Janssen, <strong>and</strong> T. Van Cauteren, “Bridg<strong>in</strong>g two ways ofdescrib<strong>in</strong>g f<strong>in</strong>al-state <strong>in</strong>teractions <strong>in</strong> A(e, e ′ p) reactions,” Phys. Lett. B 527, 62 (2002).[58] D. Debruyne <strong>and</strong> J. Ryckebusch, “A(e, e ′ p) reactions at GeV energies,” Nucl. Phys. A 699,65 (2002).[59] J. Ryckebusch, D. Debruyne, P. Lava, S. Janssen, B. Van Overmeire, <strong>and</strong> T. Van Cauteren,“Relativistic formulation of Glauber theory for A(e, e ′ p) reactions,” Nucl. Phys. A 728,226 (2003).[60] A. Sommerfeld <strong>and</strong> J. Runge, “Anwendung der Vektorrechnung auf die Grundlagen dergeometrischen Optik,” Ann. Physik 340, 277 (1911).[61] M. Born <strong>and</strong> E. Wolf, Pr<strong>in</strong>ciples of Optics (Pergamon Press, Oxford, 1980).[62] R. J. Glauber <strong>and</strong> G. Matthiae, “High-energy scatter<strong>in</strong>g of protons by nuclei,” Nucl. Phys.B 21, 135 (1970).[63] S. J. Wallace, “High-energy expansion for nuclear multiple scatter<strong>in</strong>g,” Phys. Rev. C 12,179 (1975).[64] D. R. Yennie, “Interaction of High-Energy Photons with Nuclei as a Test of Vector-Meson-Dom<strong>in</strong>ance,” <strong>in</strong> Hadronic Interactions of Electrons <strong>and</strong> Photons, J. Cumm<strong>in</strong>gs <strong>and</strong> D. Osborn,eds., (Academic Press, New York, 1971), p. 321.[65] J. D. Bjorken <strong>and</strong> S. D. Drell, Relativistic Quantum Mechanics (McGraw-Hill, New York,1964).[66] J. A. Caballero, T. W. Donnelly, E. Moya de Guerra, <strong>and</strong> J. M. Udías, “Analysis of factorization<strong>in</strong> (e, e ′ p) reactions: A survey of the relativistic plane wave impulse approximation,”Nucl. Phys. A 632, 323 (1998).[67] I. Yaron, L. Frankfurt, E. Piasetzky, M. Sargsian, <strong>and</strong> M. Strikman, “Investigation ofthe high momentum components of the nuclear wave function us<strong>in</strong>g hard quasielasticA(p, 2p)X reactions,” Phys. Rev. C 66, 024601 (2002).[68] G. D. Alkhazov, S. L. Belostotsky, <strong>and</strong> A. A. Voroboyov, “Scatter<strong>in</strong>g of 1 GeV protons onnuclei,” Phys. Rep. 42, 89 (1978).[69] J. R. Vignote, M. C. Martínez, J. A. Caballero, E. Moya de Guerra, <strong>and</strong> J. M. Udías,“A(⃗e, e ′ ⃗p)B responses: From bare nucleons to complex nuclei,” Phys. Rev. C 70, 044608(2004).[70] R. D. Amado, J. Piekarewicz, D. A. Sparrow, <strong>and</strong> J. A. McNeil, “Dirac-eikonal scatter<strong>in</strong>gamplitude,” Phys. Rev. C 28, 1663 (1983).


Bibliography 106[71] R. J. Glauber, <strong>in</strong> Lectures <strong>in</strong> <strong>Theoretical</strong> <strong>Physics</strong>, W. E. Britt<strong>in</strong> <strong>and</strong> L. G. Bunham, eds., (Interscience,New York, 1959), Vol. 1, p. 315.[72] W. R. Greenberg <strong>and</strong> G. A. Miller, “Color transparency <strong>and</strong> Dirac-based sp<strong>in</strong> effects <strong>in</strong>(e, e ′ p) reactions,” Phys. Rev. C 49, 2747 (1994).[73] H. Ito, S. E. Koon<strong>in</strong>, <strong>and</strong> R. Seki, “Sp<strong>in</strong>-polarization response functions <strong>in</strong> high-energy(⃗e, e ′ ⃗p) reactions,” Phys. Rev. C 56, 3231 (1997).[74] A. Bianconi <strong>and</strong> M. Radici, “A test of the eikonal approximation <strong>in</strong> high-energy (e, e ′ p)scatter<strong>in</strong>g,” Phys. Lett. B 363, 24 (1995).[75] A. Bianconi <strong>and</strong> M. Radici, “Angular distributions for knockout <strong>and</strong> scatter<strong>in</strong>g of protons<strong>in</strong> the eikonal approximation,” Phys. Rev. C 54, 3117 (1996).[76] M. Radici, W. H. Dickhoff, <strong>and</strong> E. R. Stoddard, “Consistency of spectroscopic factors from(e, e ′ p) reactions at different momentum transfers,” Phys. Rev. C 66, 014613 (2002).[77] G. R. Satchler, Direct <strong>Nuclear</strong> Reactions (Oxford University Press, Oxford, 1983).[78] J. J. Kelly, “Effects of sp<strong>in</strong>or distortion <strong>and</strong> density-dependent form factors uponquasifree 16 O(⃗e, e ′ ⃗p),” Phys. Rev. C 60, 044609 (1999).[79] R. H. Bassel <strong>and</strong> C. Wilk<strong>in</strong>, “High-Energy Proton Scatter<strong>in</strong>g <strong>and</strong> the Structure of LightNuclei,” Phys. Rev. 174, 1179 (1968).[80] L. L. Frankfurt, E. Moniz, M. Sargsyan, <strong>and</strong> M. I. Strikman, “Correlation effects <strong>in</strong> nucleartransparency,” Phys. Rev. C 51, 3435 (1995).[81] N. N. Nikolaev, A. Szcurek, J. Speth, J. Wambach, B. G. Zakharov, <strong>and</strong> V. R. Zoller,“Multiple-scatter<strong>in</strong>g effects <strong>in</strong> the transverse-momentum distributions from (e, e ′ p) reactions,”Nucl. Phys. A 582, 665 (1995).[82] S. Jeschonnek <strong>and</strong> T. W. Donnelly, “Sp<strong>in</strong>-orbit f<strong>in</strong>al-state <strong>in</strong>teraction <strong>in</strong> the framework ofGlauber theory for (e, e ′ p) reactions,” Phys. Rev. C 59, 2676 (1999).[83] O. Benhar, N. Nikolaev, J. Speth, A. Usmani, <strong>and</strong> B. Zakharov, “F<strong>in</strong>al state <strong>in</strong>teractions <strong>in</strong>4 He(e, e ′ p) 3 H at large proton energy,” Nucl. Phys. A 673, 241 (2000).[84] A. Kohama, K. Yazaki, <strong>and</strong> R. Seki, “Correlation <strong>and</strong> f<strong>in</strong>ite <strong>in</strong>teraction-range effects <strong>in</strong>high-energy electron <strong>in</strong>clusive scatter<strong>in</strong>g,” Nucl. Phys. A 662, 175 (2000).[85] C. Ciofi degli Atti, L. P. Kaptari, <strong>and</strong> D. Treleani, “Effects of the f<strong>in</strong>al state <strong>in</strong>teraction <strong>in</strong>the electrodis<strong>in</strong>tegration of the deuteron at <strong>in</strong>termediate <strong>and</strong> high energies,” Phys. Rev.C 63, 044601 (2001).


Bibliography 107[86] M. Petraki, E. Mavrommatis, O. Benhar, J. W. Clark, A. Fabroc<strong>in</strong>i, <strong>and</strong> S. Fantoni, “F<strong>in</strong>alstate<strong>in</strong>teractions <strong>in</strong> the response of nuclear matter,” Phys. Rev. C 67, 014605 (2003).[87] P. Lava, M. C. Martínez, J. Ryckebusch, J. A. Caballero, <strong>and</strong> J. M. Udías, “<strong>Nuclear</strong> transparencies<strong>in</strong> relativistic A(e, e ′ p) models,” Phys. Lett. B 595, 177 (2004).[88] P. Lava, J. Ryckebusch, B. Van Overmeire, <strong>and</strong> S. Strauch, “Polarization transfer <strong>in</strong>4 He(⃗e, e ′ ⃗p) <strong>and</strong> 16 O(⃗e, e ′ ⃗p) <strong>in</strong> a relativistic Glauber model,” Phys. Rev. C 71, 014605 (2005).[89] C. Lechano<strong>in</strong>e-LeLuc <strong>and</strong> F. Lehar, “Nucleon-nucleon elastic scatter<strong>in</strong>g <strong>and</strong> total crosssections,” Rev. Mod. Phys. 65, 47 (1993).[90] S. J. Wallace, “High-Energy Proton Scatter<strong>in</strong>g,” Adv. Nucl. Phys. 12, 135 (1981).[91] A. D. Krisch, “The Sp<strong>in</strong> of the Proton,” Scientific American 240, 58 (1979).[92] F. Bauer et al., “Measurement of Sp<strong>in</strong> Correlation Parameters A NN , A SS , <strong>and</strong> A SL at2.1 GeV <strong>in</strong> Proton-Proton Elastic Scatter<strong>in</strong>g,” Phys. Rev. Lett. 90, 142301 (2003).[93] K. O. Eyser, R. Machleidt, W. Scobel, <strong>and</strong> the EDDA Collaboration, “Modell<strong>in</strong>g nucleonnucleonscatter<strong>in</strong>g above 1 GeV,” Eur. Phys. J. A 22, 105 (2004).[94] K. Hagiwara et al., “Review of Particle <strong>Physics</strong>,” Phys. Rev. D 66, 010001 (2002).[95] Particle Data Group, http://pdg.lbl.gov .[96] A. Dobrovolsky et al., “Small angle pp scatter<strong>in</strong>g at energies from 650 to 1000 MeV,” Nucl.Phys. B 214, 1 (1983).[97] B. Silverman et al., “Differential cross sections <strong>and</strong> analyz<strong>in</strong>g powers for small angleneutron-proton elastic scatter<strong>in</strong>g between 378 <strong>and</strong> 1135 MeV,” Nucl. Phys. A 499, 763(1989).[98] R. E. Hendrick <strong>and</strong> B. Lautrup, “Real parts of the forward elastic π ± p, K ± p, ¯pp, <strong>and</strong> ppscatter<strong>in</strong>g amplitudes from 1 to 200 GeV/c,” Phys. Rev. D 11, 529 (1975).[99] A. A. Carter <strong>and</strong> D. V. Bugg, “The real part of the proton-neutron forward scatter<strong>in</strong>gamplitude,” Phys. Lett. 20, 203 (1966).[100] C. J. Joacha<strong>in</strong> <strong>and</strong> C. Quigg, “Multiple scatter<strong>in</strong>g expansions <strong>in</strong> several particele dynamics,”Rev. Mod. Phys. 46, 279 (1974).[101] S. J. Wallace, “Eikonal Expansion,” Phys. Rev. Lett. 27, 622 (1971).[102] S. J. Wallace, “Eikonal expansion,” Ann. Phys. (N.Y.) 78, 190 (1973).[103] S. J. Wallace, “High-Energy Expansions of Scatter<strong>in</strong>g Amplitudes,” Phys. Rev. D 8, 1846(1973).


Bibliography 108[104] S. J. Wallace <strong>and</strong> J. A. McNeil, “Relativistic eikonal expansion,” Phys. Rev. D 16, 3565(1977).[105] S. J. Wallace <strong>and</strong> J. L. Friar, “Approximate Dirac scatter<strong>in</strong>g amplitudes: Eikonal expansion,”Phys. Rev. C 29, 956 (1984).[106] D. Waxman, C. Wilk<strong>in</strong>, J.-F. Germond, <strong>and</strong> R. J. Lombard, “Eikonal corrections for sp<strong>in</strong>orbitpotentials,” Phys. Rev. C 24, 578 (1981).[107] G. Fäldt, A. Ingemarsson, <strong>and</strong> J. Mahalanabis, “Numerical <strong>in</strong>vestigation of non-eikonalcorrections to the Glauber model at <strong>in</strong>termediate energies,” Phys. Rev. C 46, 1974 (1992).[108] F. Carstoiu <strong>and</strong> R. J. Lombard, “Eikonal expansion for total cross sections of heavy ionreactions,” Phys. Rev. C 48, 830 (1993).[109] M. H. Cha <strong>and</strong> Y. J. Kim, “Higher-order corrections to the eikonal phase shifts for heavyionelastic collision,” Phys. Rev. C 51, 212 (1995).[110] J. S. Al-Khalili, J. A. Tostev<strong>in</strong>, <strong>and</strong> J. M. Brooke, “Beyond the eikonal model for few-bodysystems,” Phys. Rev. C 55, R1018 (1997).[111] C. E. Aguiar, F. Zardi, <strong>and</strong> A. Vitturi, “Low-energy extensions of the eikonal approximationto heavy-ion scatter<strong>in</strong>g,” Phys. Rev. C 56, 1511 (1997).[112] J. A. Tjon <strong>and</strong> S. J. Wallace, “Coulomb corrections <strong>in</strong> quasielastic scatter<strong>in</strong>g based on theeikonal expansion for electron wave functions,” Phys. Rev. C 74, 064602 (2006).[113] A. Baker, “Second-Order Eikonal Approximation for Potential Scatter<strong>in</strong>g,” Phys. Rev. D6, 3462 (1972).[114] S. L. Belostotsky et al., “Quasielastic proton scatter<strong>in</strong>g at 1.0 GeV,” In Proceed<strong>in</strong>gs of theInternational Symposium on Modern Developments <strong>in</strong> <strong>Nuclear</strong> <strong>Physics</strong>, p. 191 (Novosibirsk,1987).[115] H. Tyrén, S. Kull<strong>and</strong>er, O. Sundberg, R. Ramach<strong>and</strong>ran, P. Isacsson, <strong>and</strong> T. Berggren,“Quasi-free proton-proton scatter<strong>in</strong>g <strong>in</strong> light nuclei at 460 MeV,” Nucl. Phys. 79, 321(1966).[116] W. J. McDonald et al., “Quasi-free nucleon scatter<strong>in</strong>g on 16 O,” Nucl. Phys. A 456, 577(1986).[117] B. Van Overmeire, W. Cosyn, P. Lava, <strong>and</strong> J. Ryckebusch, “Relativistic eikonal descriptionof A(p, pN) reactions,” Phys. Rev. C 73, 064603 (2006).[118] K. Hatanaka et al., “Exclusive Measurement of Proton Quasifree Scatter<strong>in</strong>g <strong>and</strong> DensityDependence of the Nucleon-Nucleon Interaction,” Phys. Rev. Lett. 78, 1014 (1997).


Bibliography 109[119] L. R. B. Elton <strong>and</strong> O. Sundberg, “<strong>Nuclear</strong> structure <strong>in</strong>formation from high-energy (p, 2p)experiments,” Nucl. Phys. A 187, 314 (1972).[120] V. P<strong>and</strong>harip<strong>and</strong>e, I. Sick, <strong>and</strong> P. K. A. deWitt Huberts, “Independent particle motion<strong>and</strong> correlations <strong>in</strong> fermion systems,” Rev. Mod. Phys. 69, 981 (1997).[121] G. J. Kramer, Ph.D. thesis, University of Amsterdam, 1990.[122] A. H. Mueller, “Topics <strong>in</strong> high energy perturbative QCD <strong>in</strong>clud<strong>in</strong>g <strong>in</strong>teractions with nuclearmatter,” In Proceed<strong>in</strong>gs of the XVII Rencontre de Moriond, Les Arcs, France, J. TranThanh Van, ed., p. 13 (Editions Frontières, Gif-sur-Yvette, 1982).[123] S. J. Brodsky, “Breakdown of QCD factorization theorems for <strong>in</strong>clusive reactions,” InProceed<strong>in</strong>gs of the XIII International Symposium on Multiparticle Dynamics, Volendam, TheNetherl<strong>and</strong>s, E. W. Kittel, W. Metzger, <strong>and</strong> A. Stergiou, eds., p. 963 (World Scientific, S<strong>in</strong>gapore,1982).[124] A. Carroll et al., “<strong>Nuclear</strong> Transparency to Large-Angle pp Elastic Scatter<strong>in</strong>g,” Phys. Rev.Lett. 61, 1698 (1988).[125] I. Mardor et al., “<strong>Nuclear</strong> Transparency <strong>in</strong> Large Momentum Transfer Quasielastic Scatter<strong>in</strong>g,”Phys. Rev. Lett. 81, 5085 (1998).[126] A. Leksanov et al., “Energy Dependence of <strong>Nuclear</strong> Transparency <strong>in</strong> C(p, 2p) Scatter<strong>in</strong>g,”Phys. Rev. Lett. 87, 212301 (2001).[127] A. W. Hendry, “Evidence for coherent effects <strong>in</strong> large-angle hadron-hadron scatter<strong>in</strong>g,”Phys. Rev. D 10, 2300 (1974).[128] D. P. Owen, F. C. Peterson, J. Orear, A. L. Read, D. G. Ryan, D. H. White, A. Ashmore,C. J. S. Damerell, W. R. Frisken, <strong>and</strong> R. Rub<strong>in</strong>ste<strong>in</strong>, “High-Energy Elastic Scatter<strong>in</strong>g of π ± ,K − , <strong>and</strong> p − on Hydrogen at c.m. Angles from 22 ◦ to 180 ◦ ,” Phys. Rev. 181, 1794 (1969).[129] K. A. Jenk<strong>in</strong>s, L. E. Price, R. Klem, R. J. Miller, P. Schre<strong>in</strong>er, H. Courant, Y. I. Makdisi, M. L.Marshak, E. A. Peterson, <strong>and</strong> K. Ruddick, “Measurement of the Energy Dependence ofElastic πp <strong>and</strong> pp Scatter<strong>in</strong>g at Large Angles,” Phys. Rev. Lett. 40, 425 (1978).[130] K. A. Jenk<strong>in</strong>s, L. E. Price, R. Klem, R. J. Miller, P. Schre<strong>in</strong>er, H. Courant, Y. I. Makdisi,M. L. Marshak, E. A. Peterson, <strong>and</strong> K. Ruddick, “Fluctuations <strong>in</strong> Large-Angle π ± p ElasticScatter<strong>in</strong>g,” Phys. Rev. Lett. 40, 429 (1978).[131] K. A. Jenk<strong>in</strong>s, L. E. Price, R. Klem, R. J. Miller, P. Schre<strong>in</strong>er, M. L. Marshak, E. A. Peterson,<strong>and</strong> K. Ruddick, “Measurement of wide-angle elastic scatter<strong>in</strong>g of pions <strong>and</strong> protons offprotons,” Phys. Rev. D 21, 2445 (1980).


Bibliography 110[132] C. Bagl<strong>in</strong> et al., “An experimental study of large-angle elastic scatter<strong>in</strong>g of chargedmesons <strong>and</strong> antiprotons on protons at 20 <strong>and</strong> 30 GeV/c <strong>in</strong>cident momenta,” Nucl. Phys.B 216, 1 (1983).[133] J. Blazey, Ph.D. thesis, University of M<strong>in</strong>nesota, 1986.[134] P. Ja<strong>in</strong>, B. Kundu, <strong>and</strong> J. P. Ralston, “Oscillat<strong>in</strong>g color transparency <strong>in</strong> πA → πp(A − 1)<strong>and</strong> γA → πN(A − 1),” Phys. Rev. D 65, 094027 (2002).[135] L. Y. Zhu et al., “Cross-Section Measurement of Charged-Pion Photoproduction from Hydrogen<strong>and</strong> Deuterium,” Phys. Rev. Lett. 91, 022003 (2003).[136] L. Y. Zhu et al., “Cross section measurements of charged pion photoproduction <strong>in</strong> hydrogen<strong>and</strong> deuterium from 1.1 to 5.5 GeV,” Phys. Rev. C 71, 044603 (2005).[137] R. L. Anderson, D. B. Gustavson, D. M. Ritson, G. A. Weitsch, H. J. Halpern, R. Prepost,D. H. Tompk<strong>in</strong>s, <strong>and</strong> D. E. Wisner, “Measurements of exclusive photoproductionprocesses at large values of t <strong>and</strong> u from 4 to 7.5 GeV,” Phys. Rev. D 14, 679 (1976).[138] J. Napolitano et al., “Measurement of the Differential Cross Section for the reaction2 H(γ, p)n at High Photon Energies <strong>and</strong> θ c.m. = 90 ◦ ,” Phys. Rev. Lett. 61, 2530 (1988).[139] S. J. Freedman et al., “Two-body dis<strong>in</strong>tegration of the deuteron with 0.8-1.8 GeV photons,”Phys. Rev. C 48, 1864 (1993).[140] J. E. Belz et al., “Two-Body Photodis<strong>in</strong>tegration of the Deuteron up to 2.8 GeV,” Phys.Rev. Lett. 74, 646 (1995).[141] C. Bochna et al., “Measurements of Deuteron Photodis<strong>in</strong>tegration up to 4.0 GeV,” Phys.Rev. Lett. 81, 4576 (1998).[142] E. C. Schulte et al., “Measurement of the High Energy Two-Body Deuteron Photodis<strong>in</strong>tegrationDifferential Cross Section,” Phys. Rev. Lett. 87, 102302 (2001).[143] L. L. Frankfurt, G. A. Miller, M. M. Sargsian, <strong>and</strong> M. I. Strikman, “QCD Rescatter<strong>in</strong>g <strong>and</strong>High Energy Two-Body Photodis<strong>in</strong>tegration of the Deuteron,” Phys. Rev. Lett. 84, 3045(2000).[144] J. P. Ralston <strong>and</strong> B. Pire, “Fluctuat<strong>in</strong>g Proton Size <strong>and</strong> Oscillat<strong>in</strong>g Color Transparency,”Phys. Rev. Lett. 61, 1823 (1988).[145] S. J. Brodsky <strong>and</strong> G. F. de Teramond, “Sp<strong>in</strong> Correlations, QCD Color Transparency, <strong>and</strong>Heavy-Quark Thresholds <strong>in</strong> Proton-Proton Scatter<strong>in</strong>g,” Phys. Rev. Lett. 60, 1924 (1988).[146] J. P. Ralston <strong>and</strong> B. Pire, “Quantum color transparency,” Phys. Rev. Lett. 65, 2343 (1990).


Bibliography 111[147] P. Ja<strong>in</strong>, B. Pire, <strong>and</strong> J. P. Ralston, “Quantum color transparency <strong>and</strong> nuclear filter<strong>in</strong>g,”Phys. Rep. 271, 67 (1996).[148] G. R. Farrar, H. Liu, L. L. Frankfurt, <strong>and</strong> M. I. Strikman, “Transparency <strong>in</strong> <strong>Nuclear</strong>Quasiexclusive Processes with Large Momentum Transfer,” Phys. Rev. Lett. 61, 686(1988).[149] B. K. Jenn<strong>in</strong>gs <strong>and</strong> G. A. Miller, “Energy dependence of color transparency,” Phys. Rev.D 44, 692 (1991).[150] B. Van Overmeire <strong>and</strong> J. Ryckebusch, “A relativistic framework to determ<strong>in</strong>e the nucleartransparency from A(p, 2p) reactions,” Phys. Lett. B 644, 304 (2007).[151] S. C. Pieper <strong>and</strong> R. B. Wir<strong>in</strong>ga, “Quantum Monte Carlo Calculations of Light Nuclei,”Ann. Rev. Nucl. Part. Sci. 51, 53 (2001).[152] P. Ja<strong>in</strong> <strong>and</strong> J. P. Ralston, “Exclusive Hadronic Processes <strong>and</strong> Color Transparency,” Pramana57, 433 (2001).[153] H. Gao, “Exclusive processes at Jefferson Lab,” Pramana 61, 837 (2003).[154] C. G. White et al., “Comparison of 20 exclusive reactions at large t,” Phys. Rev. D 49, 58(1994).[155] S. J. Brodsky <strong>and</strong> G. R. Farrar, “Scal<strong>in</strong>g Laws at Large Transverse Momentum,” Phys.Rev. Lett. 31, 1153 (1973).[156] S. J. Brodsky <strong>and</strong> G. R. Farrar, “Scal<strong>in</strong>g laws for large-momentum-transfer processes,”Phys. Rev. D 11, 1309 (1975).[157] V. A. Matveev, R. M. Muradian, <strong>and</strong> A. N. Tavkhelidze, “Automodellism <strong>in</strong> the largeangleelastic scatter<strong>in</strong>g <strong>and</strong> structure of hadrons,” Lett. Nuovo Cimento 7, 719 (1973).[158] G. P. Lepage <strong>and</strong> S. J. Brodsky, “Exclusive processes <strong>in</strong> perturbative quantum chromodynamics,”Phys. Rev. D 22, 2157 (1980).[159] D. Sivers, S. J. Brodsky, <strong>and</strong> R. Blankenbecler, “Large transverse momentum processes,”Phys. Rep. 23, 1 (1976).[160] G. C. Blazey et al., “Hard Scatter<strong>in</strong>g with Exclusive Reactions: π − p Elastic Scatter<strong>in</strong>g <strong>and</strong>ρ-Meson Production,” Phys. Rev. Lett. 55, 1820 (1985).[161] B. R. Baller et al., “Comparison of exclusive reactions at large t,” Phys. Rev. Lett. 60, 1118(1988).[162] S. J. Brodsky <strong>and</strong> G. P. Lepage, “Helicity selection rules <strong>and</strong> tests of gluon sp<strong>in</strong> <strong>in</strong> exclusivequantum-chromodynamic processes,” Phys. Rev. D 24, 2848 (1981).


Bibliography 112[163] K. Wijesooriya et al., “Polarization Measurements <strong>in</strong> High-Energy Deuteron Photodis<strong>in</strong>tegration,”Phys. Rev. Lett. 86, 2975 (2001).[164] K. Wijesooriya et al., “Polarization measurements <strong>in</strong> neutral pion photoproduction,”Phys. Rev. C 66, 034614 (2002).[165] D. Abbott et al., “Measurement of Tensor Polarization <strong>in</strong> Elastic Electron-Deuteron Scatter<strong>in</strong>gat Large Momentum Transfer,” Phys. Rev. Lett. 84, 5053 (2000).[166] F. Adamian et al., “Measurement of the cross-section asymmetry of deuteron photodis<strong>in</strong>tegrationprocess by l<strong>in</strong>early polarized photons <strong>in</strong> the energy range E γ = 0.8–1.6 GeV,”Eur. Phys. J. A 8, 423 (2000).[167] T. Gousset, B. Pire, <strong>and</strong> J. P. Ralston, “Hadron helicity violation <strong>in</strong> exclusive processes:Quantitative calculations <strong>in</strong> lead<strong>in</strong>g order QCD,” Phys. Rev. D 53, 1202 (1996).[168] J. P. Ralston <strong>and</strong> P. Ja<strong>in</strong>, “QCD Form Factors <strong>and</strong> Hadron Helicity Non-Conservation,”Phys. Rev. D 69, 053008 (2004).[169] D. G. Crabb et al., “Sp<strong>in</strong> Dependence of High-P⊥ 2 Elastic p−p Scatter<strong>in</strong>g,” Phys. Rev. Lett.41, 1257 (1978).[170] E. A. Crosbie et al., “Energy dependence of sp<strong>in</strong>-sp<strong>in</strong> effects <strong>in</strong> p − p elastic scatter<strong>in</strong>g at90 ◦ c.m.,” Phys. Rev. D 23, 600 (1981).[171] S. L. L<strong>in</strong>n et al., “Measurements of sp<strong>in</strong> parameters <strong>in</strong> p − p elastic scatter<strong>in</strong>g at 6 GeV/c,”Phys. Rev. D 26, 550 (1982).[172] T. S. Bhatia et al., “Sp<strong>in</strong> Correlation for pp Elastic Scatter<strong>in</strong>g at θ c.m. = π/2 <strong>in</strong> the EnergyRegion of Dibaryon Resonances,” Phys. Rev. Lett. 49, 1135 (1982).[173] G. R. Court et al., “Energy Dependence of Sp<strong>in</strong> Effects <strong>in</strong> p ↑ +p ↑ → p+p,” Phys. Rev. Lett.57, 507 (1986).[174] N. Isgur <strong>and</strong> C. H. Llewellyn Smith, “Asymptotic Q 2 for Exclusive Processes <strong>in</strong> QuantumChromodynamics,” Phys. Rev. Lett. 52, 1080 (1984).[175] A. V. Radyushk<strong>in</strong>, “Sum Rules And Exclusive Processes In QCD,” Acta Phys. Polon. B15, 403 (1984).[176] A. P. Bakulev <strong>and</strong> A. V. Radyushk<strong>in</strong>, “Nonlocal condensates <strong>and</strong> QCD sum rules for thepion form-factor,” Phys. Lett. B 271, 223 (1991).[177] G. Sterman <strong>and</strong> P. Stoler, “Hadronic Form Factors <strong>and</strong> Perturbative QCD,” Ann. Rev.Nucl. Part. Sci. 47, 193 (1997).


Bibliography 113[178] B. Pire <strong>and</strong> J. P. Ralston, “Fixed angle elastic scatter<strong>in</strong>g <strong>and</strong> the chromo-Coulomb phaseshift,” Phys. Lett. B 117, 233 (1982).[179] P. V. L<strong>and</strong>shoff <strong>and</strong> J. C. Polk<strong>in</strong>ghorne, “Models for hadronic <strong>and</strong> leptonic processes athigh energy,” Phys. Rep. 5, 1 (1972).[180] P. V. L<strong>and</strong>shoff, “Model for elastic scatter<strong>in</strong>g at wide angle,” Phys. Rev. D 10, 1024 (1974).[181] P. V. L<strong>and</strong>shoff <strong>and</strong> D. J. Pritchard, “Higher Order QCD Corrections to Exclusive Processes:the Multiple Scatter<strong>in</strong>g Mechanism,” Z. Phys. C 6, 69 (1980).[182] J. C. Polk<strong>in</strong>ghorne, “Quark-Quark Interactions <strong>and</strong> High Energy Processes,” Phys. Lett.B 49, 277 (1974).[183] V. V. Sudakov, “Vertex parts at very high energies <strong>in</strong> quantum electrodynamics,” Sov.Phys. JETP 3, 65 (1956).[184] A. H. Mueller, “Perturbative QCD at high energies,” Phys. Rep. 73, 237 (1981).[185] J. Botts <strong>and</strong> G. Sterman, “Sudakov effects <strong>in</strong> hadron-hadron elastic scatter<strong>in</strong>g,” Phys. Lett.B 224, 201 (1989).[186] J. Botts <strong>and</strong> G. Sterman, “Sudakov effects <strong>in</strong> hadron-hadron elastic scatter<strong>in</strong>g: J. Botts <strong>and</strong>G. Sterman, Phys. Lett. B 224 (1989) 201,” Phys. Lett. B 227, 501(E) (1989).[187] J. P. Ralston <strong>and</strong> B. Pire, “Oscillatory Scal<strong>in</strong>g Violations <strong>and</strong> the Quantum ChromodynamicCoulomb Phase,” Phys. Rev. Lett. 49, 1605 (1982).[188] A. Sen, “Asymptotic behavior of the fixed-angle on-shell quark scatter<strong>in</strong>g amplitudes <strong>in</strong>non-Abelian gauge theories,” Phys. Rev. D 28, 860 (1983).[189] G. B. West <strong>and</strong> D. R. Yennie, “Coulomb Interference <strong>in</strong> High-Energy Scatter<strong>in</strong>g,” Phys.Rev. 172, 1413 (1968).[190] V. Franco, “Coulomb-<strong>Nuclear</strong> Interference,” Phys. Rev. D 7, 215 (1973).[191] D. Zwanziger, “Reduction Formulas for Charged Particles <strong>and</strong> Coherent States <strong>in</strong> QuantumElectrodynamics,” Phys. Rev. D 7, 1082 (1973).[192] L. L. Frankfurt, E. Piasetsky, M. M. Sargsyan, <strong>and</strong> M. I. Strikman, “Prob<strong>in</strong>g short-rangenucleon correlations <strong>in</strong> high-energy hard quasielastic pd reactions,” Phys. Rev. C 51, 890(1995).[193] R. J. Furnstahl, B. D. Serot, <strong>and</strong> H.-B. Tang, “A chiral effective lagrangian for nuclei,”Nucl. Phys. A 615, 441 (1997).[194] L. L. Frankfurt <strong>and</strong> M. I. Strikman, “High-energy phenomena, short-range nuclear structure<strong>and</strong> QCD,” Phys. Rep. 76, 215 (1981).


Bibliography 114[195] L. L. Frankfurt <strong>and</strong> M. I. Strikman, “Color Screen<strong>in</strong>g <strong>and</strong> Color Transparency <strong>in</strong> Hard<strong>Nuclear</strong> Processes,” Prog. Part. Nucl. Phys. 27, 135 (1991).[196] L. L. Frankfurt, W. R. Greenberg, G. A. Miller, <strong>and</strong> M. I. Strikman, “Sum rule descriptionof color transparency,” Phys. Rev. C 46, 2547 (1992).[197] L. L. Frankfurt, G. A. Miller, <strong>and</strong> M. I. Strikman, “The Geometrical Color Optics Of CoherentHigh Energy Processes,” Annu. Rev. Nucl. Part. Sci. 44, 501 (1994).[198] F. E. Low, “Model of the bare Pomeron,” Phys. Rev. D 12, 163 (1975).[199] S. Nuss<strong>in</strong>ov, “Colored-Quark Version of Some Hadronic Puzzles,” Phys. Rev. Lett. 34,1286 (1975).[200] J. F. Gunion <strong>and</strong> D. E. Soper, “Quark-count<strong>in</strong>g <strong>and</strong> hadron-size effects for total crosssections,” Phys. Rev. D 15, 2617 (1977).[201] S. J. Brodsky <strong>and</strong> A. H. Mueller, “Us<strong>in</strong>g nuclei to probe hadronization <strong>in</strong> QCD,” Phys.Lett. B 206, 685 (1988).[202] L. Frankfurt, G. A. Miller, <strong>and</strong> M. Strikman, “Color transparency phenomenon <strong>and</strong> nuclearphysics,” Comm. Nucl. Part. Phys. 21, 1 (1992).[203] L. L. Frankfurt <strong>and</strong> M. I. Strikman, “Hard nuclear processes <strong>and</strong> microscopic nuclearstructure,” Phys. Rep. 160, 235 (1988).[204] B. K. Jenn<strong>in</strong>gs <strong>and</strong> G. A. Miller, “Realistic Hadronic Matrix Element Approach to ColorTransparency,” Phys. Rev. Lett. 69, 3619 (1992).[205] L. L. Frankfurt <strong>and</strong> M. I. Strikman, “Po<strong>in</strong>t-like configurations <strong>in</strong> hadrons <strong>and</strong> nuclei <strong>and</strong>deep <strong>in</strong>elastic reactions with leptons: EMC <strong>and</strong> EMC-like effects,” Nucl. Phys. B 250, 143(1985).[206] I. Mardor, Ph.D. thesis, Tel-Aviv University, 1997.[207] J. Acl<strong>and</strong>er et al., “<strong>Nuclear</strong> transparency <strong>in</strong> 90 ◦ c.m. quasielastic A(p, 2p) reactions,” Phys.Rev. C 70, 015208 (2004).[208] C. Ciofi degli Atti <strong>and</strong> S. Simula, “Realistic model of the nucleon spectral function <strong>in</strong> few<strong>and</strong>many-nucleon systems,” Phys. Rev. C 53, 1689 (1996).[209] T.-S. H. Lee <strong>and</strong> G. A. Miller, “Color transparency <strong>and</strong> high-energy (p, 2p) nuclear reactions,”Phys. Rev. C 45, 1863 (1992).[210] B. K. Jenn<strong>in</strong>gs <strong>and</strong> G. A. Miller, “Color transparency <strong>and</strong> non-perturbative contributionsto high energy (p, pp) reactions,” Phys. Lett. B 274, 442 (1992).


Bibliography 115[211] S. Frankel, W. Frati, <strong>and</strong> N. R. Walet, “Extract<strong>in</strong>g nuclear transparency from p, 2p − A <strong>and</strong>e, e ′ p − A cross sections,” Nucl. Phys. A 580, 595 (1994).[212] B. K. Jenn<strong>in</strong>gs <strong>and</strong> G. A. Miller, “Color transparency <strong>in</strong> (p, pp) reactions,” Phys. Lett. B318, 7 (1993).[213] P. Ja<strong>in</strong> <strong>and</strong> J. P. Ralston, “Systematic analysis method for color transparency experiments,”Phys. Rev. D 48, 1104 (1993).[214] N. C. R. Mak<strong>in</strong>s et al., “Momentum transfer dependence of nuclear transparency fromthe 12 C(e, e ′ p) reaction,” Phys. Rev. Lett. 72, 1986 (1994).[215] T. G. O’Neill et al., “A-dependence of nuclear transparency <strong>in</strong> quasielastic A(e, e ′ p) athigh Q 2 ,” Phys. Lett. B 351, 87 (1995).[216] D. Abbott et al., “Quasifree (e, e ′ p) Reactions <strong>and</strong> Proton Propagation <strong>in</strong> Nuclei,” Phys.Rev. Lett. 80, 5072 (1998).[217] K. Garrow et al., “<strong>Nuclear</strong> transparency from quasielastic A(e, e ′ p) reactions up to Q 2 =8.1 (GeV/c) 2 ,” Phys. Rev. C 66, 044613 (2002).[218] D. Dutta et al., “Quasielastic (e, e ′ p) reaction on 12 C, 56 Fe, <strong>and</strong> 197 Au,” Phys. Rev. C 68,064603 (2003).[219] N. N. Nikolaev, A. Szczurek, J. Speth, J. Wambach, B. G. Zakharov, <strong>and</strong> V. R. Zoller,“<strong>Theoretical</strong> <strong>in</strong>terpretation of the NE18 experiment on nuclear transparency <strong>in</strong> A(e, e ′ p)scatter<strong>in</strong>g,” Phys. Rev. C 50, R1296 (1994).[220] L. L. Frankfurt, M. I. Strikman, <strong>and</strong> M. B. Zhalov, “Pitfalls <strong>in</strong> look<strong>in</strong>g for color transparencyat <strong>in</strong>termediate energies,” Phys. Rev. C. 50, 2189 (1994).[221] P. Lava, Ph.D. thesis, Ghent University, 2006.[222] G. Gar<strong>in</strong>o et al., “Proton propagation <strong>in</strong> nuclei studied <strong>in</strong> the (e, e ′ p) reaction,” Phys. Rev.C 45, 780 (1992).[223] D. Rohe et al., “<strong>Nuclear</strong> transparency from quasielastic 12 C(e, e ′ p),” Phys. Rev. C 72,054602 (2005).[224] S. Heppelmann, “Color transparency <strong>and</strong> nuclear filter<strong>in</strong>g,” Nucl. Phys. B Proc. Suppl.12, 159 (1990).[225] J. Botts <strong>and</strong> G. Sterman, “Hard elastic scatter<strong>in</strong>g <strong>in</strong> QCD: Lead<strong>in</strong>g behavior,” Nucl. Phys.B 325, 62 (1989).[226] J. Botts, “The multiple scatter<strong>in</strong>g contribution to fixed-angle elastic hadron-hadron scatter<strong>in</strong>g,”Nucl. Phys. B 353, 20 (1991).


Bibliography 116[227] J. Botts, J.-W. Qiu, <strong>and</strong> G. Sterman, “Elastic amplitudes <strong>and</strong> power corrections <strong>in</strong> QCD,”Nucl. Phys. A 527, 577 (1991).[228] D. Dutta <strong>and</strong> H. Gao, “Generalized count<strong>in</strong>g rule <strong>and</strong> oscillatory scal<strong>in</strong>g,” Phys. Rev. C71, 032201(R) (2005).[229] J-PARC home page, http://j-parc.jp/<strong>in</strong>dex-e.html .[230] S. Kumano, “Hadron <strong>Physics</strong> at J-PARC,” hep-ph/0608105 .[231] A. Ivanilov, private communication.[232] J. J. Kelly, “Nucleon Knockout by Intermediate Energy Electrons,” Adv. Nucl. Phys. 23,75 (1996).[233] T. de Forest, “Off-shell electron-nucleon cross sections: The impulse approximation,”Nucl. Phys. A 392, 232 (1983).[234] J. M. Udías, P. Sarriguren, E. Moya de Guerra, E. Garrido, <strong>and</strong> J. A. Caballero, “Spectroscopicfactors <strong>in</strong> 40 Ca <strong>and</strong> 208 Pb from (e, e ′ p): Fully relativistic analysis,” Phys. Rev. C 48,2731 (1993).[235] R. J. Woo et al., “Measurement of the Induced Proton Polarization P n <strong>in</strong> the 12 C(e, e ′ ⃗p)Reaction,” Phys. Rev. Lett. 80, 456 (1998).[236] J. M. Udías <strong>and</strong> J. R. Vignote, “Relativistic nuclear structure effects <strong>in</strong> (e, e ′ ⃗p),” Phys. Rev.C 62, 034302 (2000).[237] J. Ryckebusch, D. Debruyne, W. Van Nespen, <strong>and</strong> S. Janssen, “Meson <strong>and</strong> isobar degreesof freedom <strong>in</strong> A(⃗e, e ′ ⃗p) reactions at 0.2 ≤ Q 2 ≤ 0.8 (GeV/c) 2 ,” Phys. Rev. C 60, 034604(1999).[238] J. Gao et al., “Dynamical Relativistic Effects <strong>in</strong> Quasielastic 1p-Shell Proton Knockoutfrom 16 O,” Phys. Rev. Lett. 84, 3265 (2000).[239] K. G. Fissum et al., “Dynamics of the quasielastic 16 O(e, e ′ p) reaction at Q 2 ≈0.8 (GeV/c) 2 ,” Phys. Rev. C 70, 034606 (2004).[240] J. J. Kelly, “Influence of the Dirac sea on proton electromagnetic knockout,” Phys. Rev. C72, 014602 (2005).[241] L. Ch<strong>in</strong>itz et al., “Separation of the Interference Response Function R LT <strong>in</strong> the16 O(e, e ′ p) 15 N Reaction,” Phys. Rev. Lett. 67, 568 (1991).[242] M. C. Martínez, P. Lava, N. Jachowicz, J. Ryckebusch, K. Vantournhout, <strong>and</strong> J. M. Udías,“Relativistic models for quasielastic neutr<strong>in</strong>o scatter<strong>in</strong>g,” Phys. Rev. C 73, 024607 (2006).


Bibliography 117[243] P. Lava, N. Jachowicz, M. C. Martínez, <strong>and</strong> J. Ryckebusch, “Nucleon helicity asymmetries<strong>in</strong> quasielastic neutr<strong>in</strong>o-nucleus <strong>in</strong>teractions,” Phys. Rev. C 73, 064605 (2006).[244] W. Cosyn, M. C. Martínez, J. Ryckebusch, <strong>and</strong> B. Van Overmeire, “<strong>Nuclear</strong> transparenciesfrom photo<strong>in</strong>duced pion production,” Phys. Rev. C 74, 062201(R) (2006).[245] B. D. Serot, “Quantum hadrodynamics,” Rep. Prog. Phys. 55, 1855 (1992).[246] P. R<strong>in</strong>g, “Relativistic mean field theory <strong>in</strong> f<strong>in</strong>ite nuclei,” Prog. Part. Nucl. Phys. 37, 193(1996).[247] B. D. Serot <strong>and</strong> J. D. Walecka, “Recent Progress <strong>in</strong> Quantum Hadrodynamics,” Int. J. Mod.Phys. E 6, 515 (1997).[248] R. J. Furnstahl <strong>and</strong> B. D. Serot, “Quantum Hadrodynamics: Evolution <strong>and</strong> Revolution,”Comm. Nucl. Part. Phys. 2, A23 (2000).[249] B. D. Serot, “A relativistic nuclear field theory with π <strong>and</strong> ρ mesons,” Phys. Lett. B 86, 146(1979).[250] R. J. Furnstahl, H.-B. Tang, <strong>and</strong> B. D. Serot, “Vacuum contributions <strong>in</strong> a chiral effectiveLagrangian for nuclei,” Phys. Rev. C 52, 1368 (1995).[251] R. J. Furnstahl, B. D. Serot, <strong>and</strong> H.-B. Tang, “Analysis of chiral mean-field models fornuclei,” Nucl. Phys. A 598, 539 (1996).[252] L. D. Miller <strong>and</strong> A. E. S. Green, “Relativistic Self-Consistent Meson Field Theory of SphericalNuclei,” Phys. Rev. C 5, 241 (1972).[253] J. D. Walecka, Electron Scatter<strong>in</strong>g for <strong>Nuclear</strong> <strong>and</strong> Nucleon Structure (Cambridge UniversityPress, Cambridge, 2001).[254] H. De Vries, C. W. De Jager, <strong>and</strong> C. De Vries, “<strong>Nuclear</strong> charge-density-distribution parametersfrom elastic electron scatter<strong>in</strong>g,” At. Data Nucl. Data Tables 36, 495 (1987).[255] J. Carlson <strong>and</strong> R. Schiavilla, “Structure <strong>and</strong> dynamics of few-nucleon systems,” Rev. Mod.Phys. 70, 743 (1998).


Bibliography 118


Nederl<strong>and</strong>stalige samenvatt<strong>in</strong>g1 Inleid<strong>in</strong>gDe ontdekk<strong>in</strong>g van de atoomkern door Rutherford <strong>in</strong> het beg<strong>in</strong> van de 20ste eeuw luidde degeboorte <strong>in</strong> van de kernfysica. De <strong>in</strong>teresse <strong>in</strong> dit deelgebied van de fysica vloeit voort uithet feit dat de kern een dicht veeldeeltjessysteem van protonen en neutronen is. Als zodaniglevert de kern ons een uniek microscopisch kader om de verschillende fundamentele krachten(de sterke, de elektromagnetische en de zwakke wiselwerk<strong>in</strong>g) te onderzoeken. Bovendienbestaat meer dan 99.9% van de massa <strong>in</strong> het universum uit nucleaire materie. Bijgevolg speeltde kernfysica een cruciale rol <strong>in</strong> het doorgronden van onze wereld, van het <strong>in</strong>f<strong>in</strong>itesimale tothet astronomische.Om <strong>in</strong>formatie te bekomen over de kernstructuur en -dynamica zijn verstrooi<strong>in</strong>gsprocessenaan kernen zeer geschikt. Leptonische probes <strong>in</strong>terageren slechts zwak met de kern en kunnenhet ganse nucleaire volume aftasten. Hadronische verstrooi<strong>in</strong>g daarentegen brengt groterewerkzame doorsnedes met zich mee en heeft dus een groot ontdekk<strong>in</strong>gspotentieel, terwijl leptonischereacties eerder gebruikt worden voor precisie-experimenten. Door de sterke <strong>in</strong>teractiemet de trefkern is de theoretische modeller<strong>in</strong>g van hadronische verstrooi<strong>in</strong>g echter veeluitdagender en tasten hadronische probes slechts het oppervlak van de kern grondig af. Hetgrootste deel van dit werk is gewijd aan hadron-geïnduceerde verstrooi<strong>in</strong>g, meer bepaald aanexclusieve proton-kern reacties <strong>in</strong> het quasi-elastisch gebied. In dit gebied verwacht men dathet projectiel de uitstoot van één enkel nucleon teweeg brengt. Dit nucleon wordt samen methet verstrooide proton gedetecteerd.2 Theoretisch model voor A(p, pN) reactiesDe modeller<strong>in</strong>g van deze exclusieve proton-geïnduceerde A(p, pN) processen valt uiteen <strong>in</strong>drie deelproblemen:• de beschrijv<strong>in</strong>g van de harde verstrooi<strong>in</strong>g die de uitstoot van een gebonden nucleonveroorzaakt,• de beschrijv<strong>in</strong>g van de structuur van de doelwit- en restkern,119


Nederl<strong>and</strong>stalige samenvatt<strong>in</strong>g 120• de beh<strong>and</strong>el<strong>in</strong>g van de zachte <strong>in</strong>itiële- en f<strong>in</strong>ale-toest<strong>and</strong>s<strong>in</strong>teracties (IFSI) die het <strong>in</strong>komenden de twee uitga<strong>and</strong>e nucleonen aangaan met het kernmedium.Wat de beh<strong>and</strong>el<strong>in</strong>g van de harde verstrooi<strong>in</strong>g betreft bestaan er twee verschillende methodes.In ons model werd uitgegaan van de zogenaamde werkzame-doorsnede-gefactoriseerdebenader<strong>in</strong>g waarbij de vrije NN werkzame doorsnede als een evenredigheidsfactor optreedt<strong>in</strong> de differentiële A(p, pN) werkzame doorsnede. Deze techniek faalt echter wanneer <strong>in</strong> debeschrijv<strong>in</strong>g van de IFSI een sp<strong>in</strong>afhankelijkheid <strong>in</strong> reken<strong>in</strong>g wordt gebracht. In dat geval biedtde ontkoppel<strong>in</strong>g van de harde NN verstrooi<strong>in</strong>g en de nucleaire dynamica op amplitudeniveaueen alternatief.Voor de beschrijv<strong>in</strong>g van de kernstructuur hebben we een beroep gedaan op het schillenmodel.Hier<strong>in</strong> veronderstelt men dat de nucleonen onafhankelijk van elkaar bewegen <strong>in</strong> eengemiddeld-veldpotentiaal die de sterke <strong>in</strong>teractie met het omr<strong>in</strong>gend kernmedium beschrijft(onafhankelijk-deeltjesmodel). De <strong>in</strong>itiële- en f<strong>in</strong>ale-toest<strong>and</strong>sgolffuncties nemen dan ook deSlater-determ<strong>in</strong>antvorm aan waarbij de relativistische gebonden-toest<strong>and</strong> één-deeltjesgolffunctiesbepaald worden <strong>in</strong> de Hartreebenader<strong>in</strong>g van het σ − ω model.Het resterende <strong>in</strong>grediënt <strong>in</strong> A(p, pN) modellen betreft de propagatie van het <strong>in</strong>vallend,het verstrooid en het uitgestoten nucleon doorheen de kern en hun <strong>in</strong>teracties met de nucleonen<strong>in</strong> de restkern. De <strong>in</strong>terpretatie van alle quasi-elastische nucleon-uitstootreacties hangtzeer sterk af van de beschikbaarheid van realistische modellen voor de beschrijv<strong>in</strong>g van dezeIFSI effecten. Met drie nucleonen onderhevig aan IFSI is dit zeker het geval voor A(p, pN)reacties en de verschillende methodes om IFSI te beh<strong>and</strong>elen worden dan ook uitvoerig besproken.In de meeste theoretische A(p, pN) studies worden de IFSI effecten berekend b<strong>in</strong>nende context van de verstoorde-golf impulsbenader<strong>in</strong>g (DWIA). Hier<strong>in</strong> worden de verstrooi<strong>in</strong>gsgolffunctiesbepaald door middel van partiële-golfontwikkel<strong>in</strong>gen van de exacte oploss<strong>in</strong>genen worden de IFSI <strong>in</strong> reken<strong>in</strong>g gebracht met behulp van nucleon-kern optische potentialen.Parametrisaties voor deze optische potentialen worden niet bepaald uit theoretische overweg<strong>in</strong>gen,maar steunen op empirische gegevens van elastische nucleon-kern verstrooi<strong>in</strong>gsprocessen.Bij nucleon k<strong>in</strong>etische energieën hoger dan 1 GeV worden deze DWIA modellen problematischomwille van drie redenen. Ten eerste worden de partiële-golfexpansies alsmaarmoeilijker te hanteren bij hogere energieën vermits het aantal partiële golven dat nodig is omconvergentie te bereiken, steeds groter wordt. Daarnaast zijn globale parametrisaties van deoptische potentialen gewoonlijk beperkt tot k<strong>in</strong>etische energieën lager dan 1 GeV. Ten derdelijken optische-potentiaalmodellen ook onnatuurlijk voor de modeller<strong>in</strong>g van de IFSI bij hogeenergieverstrooi<strong>in</strong>gwegens het sterk <strong>in</strong>elastisch, absorberend en diffractief karakter van deonderliggende elementaire nucleon-nucleon <strong>in</strong>teractie.Een alternatieve beschrijv<strong>in</strong>g van de IFSI wordt geleverd door de eikonale benader<strong>in</strong>g.Deze benader<strong>in</strong>g v<strong>in</strong>dt haar oorsprong <strong>in</strong> de optica en behoort tot de groep van semi-klassiekebenader<strong>in</strong>gstechnieken die nuttig worden wanneer de de Broglie golflengte van het deeltjezeer kle<strong>in</strong> is <strong>in</strong> vergelijk<strong>in</strong>g met de dracht van de potentiaal waaraan het deeltje onderhevig is.


121Daarnaast wordt verondersteld dat de sterkte van de potentiaal veel kle<strong>in</strong>er is dan de energievan het deeltje. Het gebruik van de eikonale methode is gerechtvaardigd voor bots<strong>in</strong>gen overkle<strong>in</strong>e hoeken en bijgevolg staat deze techniek ook bekend als de kle<strong>in</strong>e-hoekbenader<strong>in</strong>g.In deze thesis werd een relativistisch A(p, pN) formalisme gebaseerd op de eikonale benader<strong>in</strong>guitgewerkt. Hierbij werden sp<strong>in</strong>effecten <strong>in</strong> de IFSI verwaarloosd om factorisatieop het niveau van de werkzame doorsnede toe te laten. Ons eikonaal model is zeer flexibelaangezien het zowel <strong>in</strong> comb<strong>in</strong>atie met relativistische optische potentialen (ROMEA) als<strong>in</strong> een Glauber veelvoudige-verstrooi<strong>in</strong>gsaanpak (RMSGA) kan gebruikt worden. De keuzemogelijkheidtussen deze twee verschillende technieken om de IFSI te beh<strong>and</strong>elen laat ons toeom A(p, pN) reacties te beschrijven <strong>in</strong> een groot energiegebied.Net als <strong>in</strong> de talloze DWIA modellen worden <strong>in</strong> het ROMEA formalisme de IFSI effectengeïmplementeerd <strong>in</strong> termen van optische potentialen. Het ROMEA formalisme steunt evenwelniet op een partiële-golfontwikkel<strong>in</strong>g, maar op de eikonale benader<strong>in</strong>g om de verstrooi<strong>in</strong>gsgolffunctiete berekenen. Hierdoor is het ROMEA formalisme beter geschikt om verstrooi<strong>in</strong>gbij hogere energieën te beschrijven.Bij k<strong>in</strong>etische energieën hoger dan 1 GeV kan de ROMEA aanpak niet langer gewettigdworden omwille van de eerder vermelde nadelen van optische potentialen. In dit hoge-energiegebiedbiedt Glaubertheorie soelaas. Deze theorie is een veeldeeltjesveralgemen<strong>in</strong>g van deeikonale benader<strong>in</strong>g en herstelt de l<strong>in</strong>k met de fundamentele nucleon-nucleon verstrooi<strong>in</strong>g viade <strong>in</strong>troductie van een profielfunctie. Verder postuleert ons RMSGA model l<strong>in</strong>eaire banen enbevroren toeschouwernucleonen. Het belangrijkste verschil tussen de ROMEA en RMSGAmodellen is dus dat het ROMEA model de IFSI ziet als nucleon-kern <strong>in</strong>teracties, terwijl hetRMSGA model uitgaat van een nucleon-nucleon beeld.3 Numerieke resultaten: de IFSI factor en A(p, pN) differentiële werkzamedoorsnedesIn onze code zit het volledige effect van de IFSI vervat <strong>in</strong> de zogenaamde IFSI factor. De eigenschappenvan deze functie werden onderzocht voor quasi-elastische protonverstrooi<strong>in</strong>g aan12 C, 16 O en 40 Ca bij een <strong>in</strong>komende protonenergie van 1 GeV. Dit komt overeen met de k<strong>in</strong>ematiekvan het PNPI experiment uitgevoerd door S. L. Belostotsky e.a. Zoals verwacht is de attenuatiehet sterkst <strong>in</strong> het <strong>in</strong>wendige van de kern en brengen zwaardere trefkernen grotere IFSIeffecten met zich mee. Verder blijken A(p, pN) reacties zich hoofdzakelijk af te spelen aan hetoppervlak van de trefkern ten gevolge van het sterk attenuerend karakter van de IFSI. In dit experiment,net als <strong>in</strong> vele <strong>and</strong>ere A(p, pN) reacties, is het uitgestoten nucleon vrij traag. Bijgevolgis de Glauber veelvoudige-verstrooi<strong>in</strong>gstheorie niet bruikbaar. De ROMEA bereken<strong>in</strong>gen daarentegengeven aanleid<strong>in</strong>g tot realistische IFSI factoren waarbij de verschillende parametrisatiesvoor de optische potentialen vergelijkbare resultaten opleveren.Vervolgens werden de voorspell<strong>in</strong>gen voor de werkzame doorsnedes vergeleken met dePNPI data. De ROMEA bereken<strong>in</strong>gen reproduceren de vormen van de gemeten differentiële


Nederl<strong>and</strong>stalige samenvatt<strong>in</strong>g 122werkzame doorsnedes vrij goed. Verder illustreert een vergelijk<strong>in</strong>g met de vlakke-golfresultatenhet tweevoudig effect van de IFSI. De IFSI zorgen niet alleen voor een attenuatie van de vlakkegolfwerkzame doorsnede, maar veroorzaken ook een shift <strong>in</strong> de impulsdistributie.De ROMEA code werd ook gebruikt om werkzame doorsnedes te berekenen voor de k<strong>in</strong>ematiekenvan <strong>and</strong>ere experimenten: 40 Ca(p, 2p) verstrooi<strong>in</strong>g bij een k<strong>in</strong>etische energie van460 MeV voor het <strong>in</strong>vallend proton, 16 O(p, 2p) en (p, pn) verstrooi<strong>in</strong>g bij 505 en 200 MeV, en4 He(p, 2p) verstrooi<strong>in</strong>g bij 250 MeV. Onze ROMEA resultaten geven een goede beschrijv<strong>in</strong>gvan de data voor de verschillende k<strong>in</strong>ematieken, waarmee aangetoond wordt dat ons model <strong>in</strong>een ruim energiegebied toepasbaar is.4 Nucleaire transparantie <strong>in</strong> A(p, 2p) reactiesDe RMSGA aanpak, die ongepast bleek voor de beschrijv<strong>in</strong>g van de experimentele werkzamedoorsnedes bij lagere energieën, werd wel gebruikt om de IFSI <strong>in</strong> reken<strong>in</strong>g te brengen bij destudie van de A(p, 2p) nucleaire transparantie <strong>in</strong> het GeV-energiegebied. De nucleaire transparantieis een maat voor de waarschijnlijkheid dat de <strong>in</strong>komende en uitga<strong>and</strong>e protonen doorheende kern propageren zonder met de toeschouwernucleonen te <strong>in</strong>terageren en is zeer geschikt<strong>in</strong> de zoektocht naar kleurtransparantie. Dit fenomeen zorgt voor een anomaal grotetransmissiewaarschijnlijkheid van de protonen doorheen het nucleaire medium doordat deprotonen naar een kle<strong>in</strong> kleurloos object fluctueren. Voor de implementatie van kleurtransparantie<strong>in</strong> onze code werden twee modellen gebruikt: het kwantum-diffusiemodel van Farrare.a. en het hadronische expansiemodel van Jenn<strong>in</strong>gs en Miller. Kleurtransparantie leidt tot eennucleaire transparantie die stijgt met toenemende energie van het <strong>in</strong>komend proton.De nucleaire-transparantiedata vertonen echter een oscillerende energie-afhankelijkheid.Om dit te verklaren werd het Ralston-Pire model voor de pp verstrooi<strong>in</strong>gsamplitude gebruikt.Volgens dit model bestaat de vrije pp verstrooi<strong>in</strong>gsamplitude uit een quark-count<strong>in</strong>g componentdie een kle<strong>in</strong> object voorstelt en een L<strong>and</strong>shoff component van normale grootte. De<strong>in</strong>terferentie tussen deze twee componenten verklaart de oscillaties <strong>in</strong> de vrije pp werkzamedoorsnede. In het nucleaire medium wordt de L<strong>and</strong>shoff term geabsorbeerd via het st<strong>and</strong>aardGlaubermechanisme, terwijl de quark-count<strong>in</strong>g term relatief we<strong>in</strong>ig geattenueerd wordt doorhet fenomeen van kleurtransparantie. Dit zogenaamde nucleaire-filter<strong>in</strong>gconcept werd <strong>in</strong> onsA(p, 2p) model <strong>in</strong>gebouwd.Voor zover ons bekend zijn de gedetailleerde RMSGA bereken<strong>in</strong>gen van de A(p, 2p) nucleairetransparantie de eerste van hun soort. De RMSGA procedure vereist de numerieke evaluatievan een <strong>in</strong>gewikkelde multidimensionale <strong>in</strong>tegraal die het effect van de bots<strong>in</strong>gen v<strong>and</strong>e <strong>in</strong>komende en uitga<strong>and</strong>e protonen met alle toeschouwernucleonen bepaalt. Hierbij treedtelk toeschouwernucleon op als verstrooi<strong>in</strong>gscentrum en wordt het voorgesteld door de corresponderenderelativistische gebonden-toest<strong>and</strong> golffunctie. Om de benodigde computertijdte verm<strong>in</strong>deren werden enkele bijkomende benader<strong>in</strong>gen <strong>in</strong>gevoerd. Zo werden de <strong>in</strong>dividuelegolffuncties van de toeschouwernucleonen vervangen door de gemiddelde proton- of neu-


123trondichtheid van de restkern. De nucleaire transparanties berekend met de volledige en debenaderde RMSGA methode zijn bijna identiek. Bijgevolg vormt de benaderde RMSGA aanpakeen volwaardig alternatief voor het computationeel langdradige, volledige RMSGA model;althans voor de bereken<strong>in</strong>g van geïntegreerde grootheden zoals de nucleaire transparantie.Onze bereken<strong>in</strong>gen reproduceren de energie-afhankelijkheid van de nucleaire-transparantiedatakwalitatief dankzij de <strong>in</strong>troductie van het concept van nucleaire filter<strong>in</strong>g. De belangrijksteconclusie van onze studie van de A(p, 2p) nucleaire transparantie is dat kleurtransparantieonontbeerlijk is om onze theoretische voorspell<strong>in</strong>gen te verhogen tot de waardes geobserveerd<strong>in</strong> de experimenten. De kwantitatieve overeenkomst met de data kan echter nog veel verbeterdworden. Het onderzoek naar de A(p, 2p) nucleaire transparantie blijft dan ook br<strong>and</strong>endactueel.5 Tweede-orde eikonale correcties <strong>in</strong> A(e, e ′ p) reactiesTen slotte hebben we de geldigheid van de eikonale benader<strong>in</strong>g bij lagere energieën onderzocht.Hiertoe hebben we een uitbreid<strong>in</strong>g van het ROMEA model ontwikkeld (SOROMEAgenaamd) die tweede-orde eikonale correcties <strong>in</strong> reken<strong>in</strong>g brengt. De ROMEA verstrooi<strong>in</strong>gsgolffunctieis immers slechts geldig tot eerste orde <strong>in</strong> 1/k, met k de impuls van het deeltje, enis dus vooral geschikt voor hoge-energieverstrooi<strong>in</strong>g.Bij dit onderzoek naar de tweede-orde eikonale correcties hebben we ons beperkt totA(e, e ′ p) observabelen. De bereken<strong>in</strong>gen van de A(e, e ′ p) nucleaire transparantie bevestigende verwachte energie-afhankelijkheid van de tweede-orde eikonale correcties: het effect neemttoe met dalende vier-impulstransfer Q 2 . De impact van de tweede-orde eikonale correcties opde A(e, e ′ p) nucleaire transparantie is evenwel zeer beperkt, zelfs bij Q 2 ≈ 0.2 (GeV/c) 2 . Vervolgenshebben we de “miss<strong>in</strong>g momentum” afhankelijkheid van enkele echt exclusieve A(e, e ′ p)observabelen, zoals de geïnduceerde normale polarisatie P n , de l<strong>in</strong>ks-rechts asymmetrie A LTen de differentiële werkzame doorsnede, onderzocht. Deze grootheden zijn naar verwacht<strong>in</strong>ggevoeliger voor details <strong>in</strong> de bereken<strong>in</strong>gen. Bij lage waarden van het “miss<strong>in</strong>g momentum”is het verschil tussen de ROMEA en SOROMEA voorspell<strong>in</strong>gen verwaarloosbaar kle<strong>in</strong>. In hetgebied van hoge “miss<strong>in</strong>g momentum” zijn de tweede-orde eikonale correcties echter significanten loopt het effect op tot 30%. Hierdoor bekomen we een betere overeenkomst met de dataen/of de relativistische DWIA bereken<strong>in</strong>gen van de Madrid-Sevilla groep.Onze numerieke bereken<strong>in</strong>gen tonen aan dat het effect van de tweede-orde eikonale correctiesop A(e, e ′ p) grootheden over het algemeen vrij beperkt is voor Q 2 ≥ 0.2 (GeV/c) 2 . Daarenbovenleidt een ver<strong>and</strong>er<strong>in</strong>g <strong>in</strong> de keuze voor de optische potentiaal, de stroomoperator of degebonden-toest<strong>and</strong> golffuncties tot vergelijkbare effecten.


Nederl<strong>and</strong>stalige samenvatt<strong>in</strong>g 1246 BesluitWe hebben een eikonaal formalisme ontwikkeld voor de modeller<strong>in</strong>g van de IFSI <strong>in</strong> A(p, pN)reacties. Bij lage energieën is het ROMEA model, dat de IFSI effecten implementeert door middelvan nucleon-kern optische potentialen, aan te raden; terwijl bij k<strong>in</strong>etische energieën hogerdan 1 GeV het RMSGA formalisme de beste keuze is. Het ROMEA model leverde voortreffelijkwerk <strong>in</strong> het reproduceren van de experimentele differentiële werkzame doorsnedes bijverschillende k<strong>in</strong>ematieken, terwijl het RMSGA formalisme gebruikt werd om de A(p, 2p) nucleairetransparantie te bestuderen.Om sp<strong>in</strong>effecten te kunnen <strong>in</strong>cluderen dient een A(p, pN) formalisme ontwikkeld te wordendat gefactoriseerd is op amplitudeniveau. Behalve sp<strong>in</strong>effecten kan eveneens het effect v<strong>and</strong>e tweede-orde eikonale correcties op A(p, pN) observabelen <strong>in</strong> de toekomst nader bekekenworden. Ten slotte verdient ook de A(p, 2p) nucleaire transparantie bijkomend onderzoek.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!