diff --git "a/data_all_eng_slimpj/shuffled/split2/finalzzhefo" "b/data_all_eng_slimpj/shuffled/split2/finalzzhefo" new file mode 100644--- /dev/null +++ "b/data_all_eng_slimpj/shuffled/split2/finalzzhefo" @@ -0,0 +1,5 @@ +{"text":"\\section{Introduction}\n\nThe results of this article strengthen the connection between invariants of Legendrian knots in standard contact $\\mathbb {R}^3$ and the $2$-variable Kauffman polynomial. Relations between the $2$-variable knot polynomials (HOMFLY-PT and Kauffman) and Legendrian knot theory were first realized in the work of Fuchs and Tabachnikov \\cite{FT} who observed, based on results of Bennequin \\cite{Be}, Franks--Morton--Williams \\cite{FW}, and Rudolph \\cite{Ru}, that these polynomials provide upper bounds on the Thurston--Bennequin number of a Legendrian knot. At that time, it was still unknown whether Legendrian knots in $\\mathbb {R}^3$ were determined up to Legendrian isotopy by their Thurston--Bennequin number, rotation number, and underlying smooth knot type (the so-called ``classical invariants'' of Legendrian knots). This question was soon resolved with the introduction of several non-classical invariants in the late 90's and early 2000's including the Legendrian contact homology algebra which is a differential graded algebra (DGA) coming from $J$-holomorphic curve theory that was constructed by Chekanov in \\cite{Chekanov02} and discovered independently by Eliashberg and Hofer, and combinatorial invariants introduced by Chekanov and Pushkar \\cite{ChP} defined by counting certain decompositions of front diagrams called normal rulings.\\footnote{Around the same time, generating family homology invariants capable of distinguishing Legendrian links with the same classical invariants were introduced by Traynor~\\cite{Tr}.} Interestingly, normal rulings were discovered independently by Fuchs in connection with augmentations of the Le\\-gendrian contact homology DGA. Moreover, Fuchs again pointed toward a connection between Legendrian invariants and topological knot invariants by conjecturing in \\cite{Fuchs} that a Legendrian knot should have a normal ruling if and only if the Kauffman polynomial estimate for the Thurston--Bennequin number is sharp. This conjecture was resolved affirmatively in \\cite{R1} by interpreting Chekanov and Pushkar's combinatorial invariants as polynomials, and showing that the ungraded ruling polynomial, $R^1_K(z)$, of a Legendrian knot $K \\subset \\mathbb {R}^3$ arises as a specialization\n\\begin{gather} \\label{eq:specialization}\nR^1_K(z) = F_K(a,z)|_{a^{-1}=0}\n\\end{gather}\nof the framed version of the Kauffman polynomial $F_K \\in \\mathbb {Z}\\big[a^{\\pm1}, z^{\\pm1}\\big]$; the specialization has the property that it is non-zero if and only if the Kauffman polynomial estimate for $\\operatorname{tb}(K)$ is sharp. An analogous result also established in~\\cite{R1} holds for the $2$-graded ruling polynomial, $R^2_K(z)$, and the HOMFLY-PT polynomial.\n\nInitially, the Legendrian invariance of the ruling polynomials, based on establishing bijections between rulings during bifurcations of the front diagram occurring in a Legendrian isotopy, was somewhat mysterious from the point of view of symplectic topology. Building on the earlier works \\cite{Fuchs, FI,Henry,NgSab, Sab}, Henry and the second author showed in~\\cite{HenryRu} that the ruling polynomials are in fact determined by the Legendrian contact homology DGA, $(\\mathcal{A}, \\partial)$, since their specializations at $z = q^{1\/2} -q^{-1\/2}$ with $q$ a prime power agree with normalized counts of augmentations of $(\\mathcal{A}, \\partial)$ to the finite field $\\mathbb{F}_q$, i.e., DGA representations from $(\\mathcal{A},\\partial)$ to $(\\mathbb{F}_q,0)$. Thus, in the ungraded case (\\ref{eq:specialization}) shows that counts of ungraded augmentations are actually topological (depending only on the underlying framed knot type of $K$), as they arise from a specialization of the Kauffman polynomial. In this article we extend this result by relating counts of higher dimensional (ungraded) representations of $(\\mathcal{A}, \\partial)$ with the $n$-colored Kauffman polynomials.\n\nTo give a statement of our main result, for $n \\geq 1$, let $\\operatorname{Rep}_1\\big(K,\\mathbb{F}_q^n\\big)$ denote the {\\it ungraded total $n$-dimensional representation number} of $K$ as defined in~\\cite{LeRu}; see Definition \\ref{def:total}. Let $F_{n,K}(a,q)$ denote the {\\it $n$-colored Kauffman polynomial} (for framed knots); see Definition~\\ref{def:Kauffman}. In Section~\\ref{sec:ncolored}, we define an {\\it ungraded $n$-colored ruling polynomial} denoted~$R^1_{n,K}(z)$.\n\n\\begin{Theorem} \\label{thm:main} For any Legendrian knot $K$ in $\\mathbb {R}^3$ with its standard contact structure and any $n \\geq 1$, there is a well-defined specialization $F_{n,K}(a,q)|_{a^{-1}=0}$, and we have\n\\[\n\\operatorname{Rep}_1\\big(K,\\mathbb{F}_q^n\\big)=R^1_{n,K}(z)=F_{n,K}(a,q)|_{a^{-1}=0}.\n\\]\n\\end{Theorem}\n\nAs an immediate consequence, we get:\n\n\\begin{Corollary}\nThe ungraded total $n$-dimensional representation number $\\operatorname{Rep}_1\\big(K,\\mathbb{F}_q^n\\big)$ depends only on the underlying framed knot type of $K$.\n\\end{Corollary}\nThe corollary is a significant strengthening of a result from~\\cite{NgR2012} that the existence of an ungraded representation of~$(\\mathcal{A}, \\partial)$ on $\\mathbb{F}_2^n$ depends only on the Thurston--Bennequin number and topological knot type of~$K$. Precisely how much of the Legendrian contact homology DGA is determined by the framed knot type of~$K$ remains an interesting question. See~\\cite{NgR2012} and~\\cite{Ng1} for some open conjectures along this line.\n\nA previous article \\cite{LeRu} establishes analogous results in the case of $2$-graded representations and the HOMFLY-PT polynomial, and in fact establishes the equality $\\operatorname{Rep}_m\\big(K,\\mathbb{F}_q^n\\big)=R^m_{n,K}(z)$ between the $m$-graded total representation numbers and $m$-graded colored ruling polynomials for all $m \\in \\mathbb {Z}_{\\geq 0}$ {\\it except for the ungraded case where $m=1$}. The $m=1$ case is more involved for a number of reasons. In the following we briefly review the argument from~\\cite{LeRu} and then outline our approach to Theorem~\\ref{thm:main}.\n\nFor $m \\neq 1$, the $n$-colored ruling polynomial is defined as a linear combination of satellite ruling polynomials of the form\n\\begin{gather} \\label{eq:mgraded}\nR^m_{n,K}(q) = \\frac{1}{c_n}\\sum_{\\beta \\in S_n} q^{\\lambda(\\beta)\/2} R^m_{S(K,\\beta)}(z)\\big|_{z = q^{1\/2}-q^{-1\/2}}, \\qquad m \\neq 1,\n\\end{gather}\nwhere \\looseness=1 $S(K,\\beta)$ is the Legendrian satellite of $K$ with a Legendrian positive permutation braid associated to $\\beta \\in S_n$. The same linear combination of HOMFLY-PT polynomials defines the $n$-colored HOMFLY-PT polynomial. In~\\cite{LeRu}, the total $n$-dimensional representation number is recovered from~(\\ref{eq:mgraded}) via a bijection between $m$-graded augmentations of $S(K,\\beta)$ and $n$-dimensional representations of the DGA of $K$ mapping a distinguished invertible generator into $B_\\beta \\subset {\\rm GL}(n, \\mathbb{F})$ where ${\\rm GL}(n,\\mathbb{F}) = \\sqcup_{\\beta \\in S_n}B_\\beta$ is the Bruhat decomposition. Thus, summing over all $\\beta \\in S_n$ corresponds to considering all $n$-dimensional representations of the DGA of~$K$ on~$\\mathbb{F}^n$.\n\nWhen $m=1$, the above bijection becomes modified in an interesting way, as augmentations of $S(K,\\beta)$ now correspond to (ungraded) representations of $(\\mathcal{A}, \\partial)$ on differential vector spaces of the form $\\big(\\mathbb{F}^n,d\\big)$ where $d$ varies over all (ungraded) upper triangular differential on $\\mathbb{F}^n$. (When $m \\neq 1$, $d =0$ is automatic for grading reasons.) The total $n$-dimensional representation number, $\\operatorname{Rep}_1\\big(K,\\mathbb{F}_q^n\\big)$, only counts representations with $d=0$, so the definition of $R^1_{n,K}$ needs to be changed to only take into account normal rulings corresponding to representations with $d=0$. This is done by replacing each $R_{S(K,\\beta)}^1$ in~(\\ref{eq:mgraded}) with the corresponding {\\it reduced ruling polynomial}~$\\widetilde{R}^1_{S(K,\\beta)}$ as introduced in~\\cite{NgR2012} that only counts normal rulings of $S(K,\\beta)$ that never pair two strands of the satellite that correspond to a single strand of $K$. Up to a technical point about the use of different diagrams in~\\cite{LeRu} and~\\cite{HenryRu} that the bulk of Section~\\ref{sec:5} is spent addressing, this leads to the equality $\\operatorname{Rep}_1\\big(K,\\mathbb{F}_q^n\\big)=R^1_{n,K}(z)$.\n\nEstablishing that $R^1_{n,K}(z)=F_{n,K}(a,q)|_{a^{-1}=0}$ requires a much more involved argument than for the case of the colored HOMFLY-PT polynomial and $R^2_{n,K}(z)$ (where the result is immediate from~\\cite{R1} and the definition). The $n$-colored Kauffman polynomial is defined by satelliting $K$ with the symmetrizer in the BMW algebra, $\\mathcal{Y}_n \\in \\operatorname{BMW}_n$. In addition to a sum over permutation braids as in the HOMFLY-PT case, $\\mathcal{Y}_n$ also has terms of a less explicit nature (though, see~\\cite{D}) involving tangle diagrams in $[0,1]\\times \\mathbb {R}$ with turn-backs, i.e., components that have both endpoints on the same boundary component of $[0,1]\\times \\mathbb {R}$. To relate $R^1_{n,K}$ and $F_{n,K}$, we use the combinatorics of normal rulings to find an inductive characterization of $R^1_{n,K}$ in terms of ordinary ruling polynomials rather than reduced ruling polynomials, and then compare this with an inductive characterization of $\\mathcal{Y}_n$ due to Heckenberger and Sch{\\\"u}ler~\\cite{Heck}.\n\nThe remainder of the article is organized as follows. In Section \\ref{sec:ncolored}, we define the ungraded $n$-colored ruling polynomial and establish an inductive characterization of it in Theorem \\ref{thm:main1}. In Section \\ref{sec:Kauffman}, we recall the definition of the colored Kauffman polynomial and prove the second equality of Theorem \\ref{thm:main} (see Theorem~\\ref{thm:main2}). Section \\ref{sec:5} reviews definitions of representation numbers from \\cite{LeRu} and then establishes the first equality of Theorem~\\ref{thm:main} (see Theorem~\\ref{thm:Rep}). In Section \\ref{sec:multicomp}, we close the article with a brief discussion of a modification of Theorem \\ref{thm:main} for the case of multi-component Legendrian links.\n\n\\section[The $n$-colored ungraded ruling polynomial]{The $\\boldsymbol{n}$-colored ungraded ruling polynomial} \\label{sec:ncolored}\n\nIn this section, after a brief review of ruling polynomials and Legendrian satellites, we define the ungraded $n$-colored ruling polynomial $R^1_{n,K}$ as a linear combination of {\\it reduced} ruling polynomials indexed by permutations $\\beta \\in S_n$. Reduced rulings, considered earlier in~\\cite{NgR2012}, form a restricted class of normal rulings of satellite links, so that it is not immediately clear how to describe $R^1_{n,K}$ in terms of ordinary ruling polynomials. For this purpose, we work in a Legendrian version of the $n$-stranded BMW algebra, $\\operatorname{BMW}_n^\\mathfrak{Leg}$, and inductively construct elements $L_n \\in \\operatorname{BMW}_n^\\mathfrak{Leg}$ that can be used to produce $R^1_{n,K}$ via (non-reduced) satellite ruling polynomials.\n\n\\subsection{Legendrian fronts and ruling polynomials}\nIn this article we consider Legendrian links and tangles in a $1$-jet space $J^1M$ where $M$ is one of~$\\mathbb {R}$,~$S^1$, or~$[0,1]$. In all cases, we can view $J^1M = T^*M\\times \\mathbb {R}$ as $M \\times \\mathbb {R}^2$, and using coordinates $(x,y,z)$ with $x \\in M$ and $y,z \\in \\mathbb {R}$ the contact form is ${\\rm d}z - y\\,{\\rm d}x$. Legendrian curves can be viewed via their {\\it front projection} $\\pi_{xz}\\colon J^1M \\rightarrow M \\times \\mathbb {R}$, $(x,y,z) \\mapsto (x,z)$ which is a collection of curves having cusp singularities and transverse double points but no vertical tangencies. The original Legendrian is recovered via $y= \\frac{{\\rm d}z}{{\\rm d}x}$, so in front diagrams (implicitly) the over-strand at a crossing is the strand with lesser slope (as the $y$-axis is oriented away from the viewer). Legendrian links have a {\\it contact framing} which is the framing given by the upward unit normal vector to the contact planes.\n\n\\begin{figure}[t] \\centering\n \\includegraphics{images\/normality.png}\n \\caption{Each closed curve of a normal ruling consists of a pair of companion paths with monotonically increasing $x$-coordinate beginning and ending at a common left and right cusp of~$\\pi_{xz}(L)$. At switches, paths from two different closed curves of~$\\rho$ meet and both turn a corner at a crossing. The normality condition requires that near switches the switching paths and their companion paths match one of the pictured configurations. At crossings that are not switches paths from two different closed curves cross transversally.} \\label{fig:normality}\n\\end{figure}\n\n\\begin{figure}[h!] \\centering\n \\begin{align*}\n \\raisebox{-.5cm}{\\includegraphics[scale=.7]{images\/mruling1.png}}-\\raisebox{-.5cm}{\\includegraphics[scale=.7]{images\/mruling2.png}} &=z\\left( \\;\\raisebox{-.5cm}{\\includegraphics[scale=.7]{images\/mruling3.png}}-\\; \\raisebox{-.5cm}{\\includegraphics[scale=.7]{images\/mruling4.png}}\\right), \\tag{R1} \\\\\n \\raisebox{-.5cm}{\\includegraphics[scale=.7]{images\/mruling5.png}}& =\\raisebox{-.5cm}{\\includegraphics[scale=.7]{images\/mruling6.png}}=0, \\tag{R2} \\\\\n \\raisebox{-.5cm}{\\includegraphics[scale=.7]{images\/unknot.png}} \\;\\sqcup K &=z^{-1}K. \\tag{R3}\n \\end{align*}\\vspace{-0.7cm}\n \\caption{The ungraded ruling polynomial skein relations.}\\label{fig:psuedo-gen}\n\\end{figure}\n\nRecall that for a Legendrian link $K \\subset J^1\\mathbb {R}$ a {\\it normal ruling} $\\rho$ of $K$ is a decomposition of the front diagram of $K$ into a collection of simple closed curves with corners at a left and right cusp and at {\\it switches} (adhering to the normality condition, see Fig.~\\ref{fig:normality}). For each $x=x_0$ where the front projection of $K$ does not have crossings or cusps, a normal ruling $\\rho$ divides the strands of~$K$ at $x=x_0$ into pairs, so that $\\rho$ can be viewed as a sequence of pairings of strands of~$K$. For a~more detailed discussion of normal rulings see for instance \\cite{Fuchs,NgR2012, R1, Sab}. The {\\it ungraded ruling polynomial} of $L$ (also called the $1$-graded ruling polynomial) is defined as\n\\[\nR_K^1(z):=\\sum_{\\rho\\in\\Gamma(K)}z^{j(\\rho)} \\in \\mathbb {Z}\\big[z^{\\pm1}\\big],\n\\] where the sum is over all normal rulings of $K$ and $j(\\rho)=\\#\\text{switches}-\\#\\text{right cusps}$. For Legendrian links in $J^1\\mathbb {R}$, the ungraded ruling polynomial of $K$ satisfies and is uniquely characterized by the skein relations in Fig.~\\ref{fig:psuedo-gen} and the normalization $R^1_{\\includegraphics[scale=.2]{images\/Unknot2}} = z^{-1}$. (See~\\cite{R1}.) The relations in Fig.~\\ref{fig:psuedo-gen} imply two additional relations that we will make use of, cf.\\ \\cite[Section~6]{R2}.\n\\begin{gather} \\label{eq:fish}\n \\text{fishtail relation:} \\ \\raisebox{-.35cm}{\\includegraphics[scale=.5]{images\/lfish.png}} =\\raisebox{-.35cm}{\\includegraphics[scale=.5]{images\/rfish.png}}=0,\n \\vspace{.5cm}\\\\\n \\text{double-crossing relation:} \\ \\raisebox{-.35cm}{\\includegraphics[scale=.5]{images\/dblesigma.png}}= \\raisebox{-.35cm}{\\includegraphics[scale=.5]{images\/id2.png}}+ z\\cdot\\raisebox{-.35cm}{\\includegraphics[scale=.5]{images\/cross.png}}- z\\cdot\\raisebox{-.35cm}{\\includegraphics[scale=.5]{images\/cusps.png}}.\\nonumber\n\\end{gather}\n\\begin{Remark} The double crossing relation can be used to show that the first relation of Fig.~\\ref{fig:psuedo-gen} also hold with right cusps. Moreover, the third relation from Fig.~\\ref{fig:psuedo-gen} is implied by the first two as long as $K \\neq \\varnothing$.\n\\end{Remark}\n\n\\subsection{A Legendrian BMW algebra}\n\nA {\\it Legendrian $n$-tangle} is a properly embedded Legendrian $\\alpha \\subset J^1[0,1]$ (i.e., compact with $\\partial \\alpha \\subset \\partial J^1[0,1]$) whose front projection agrees with the collection of horizontal lines $z=i$, $1\\leq i \\leq n$, near $\\partial J^1[0,1]$. Legendrian isotopies of $n$-tangles are required to remain fixed in a~neighborhood of $\\partial J^1[0,1]$. At $x=0$ and $x=1$ we enumerate the endpoints and strands of a Legendrian tangle from $1$ to $n$ with {\\it descending} $z$-coordinate. For any permutation $\\beta \\in S_n$, there is a corresponding {\\it positive permutation braid} that is a Legendrian $n$-tangle, also denoted $\\beta \\subset J^1[0,1]$, that connects endpoint $i$ at $x=0$ to endpoint $\\beta(i)$ at $x=1$ for $1 \\leq i \\leq n$. Up to Legendrian isotopy, $\\beta$ is uniquely characterized by requiring that\n\\begin{itemize}\\itemsep=0pt\n\\item[(i)] the front projection does not have cusps and,\n\\item[(ii)] for $i \\beta(j)$).\n\\end{itemize}\nThe number of crossings in such a front diagram for $\\beta$ is called the {\\it length} of $\\beta$ and will be deno\\-ted~$\\lambda(\\beta)$. For Legendrian $n$-tangles $\\alpha,\\beta \\subset J^1[0,1]$, we define their multiplication $\\alpha\\cdot\\beta \\subset J^1[0,1]$ by stacking $\\beta$ to the {\\it left} of $\\alpha$ (as in composition of permutations). Diagrammatically:\n\\[\n\\labellist\n\\small\n\\pinlabel $\\alpha$ at 45 38\n\\endlabellist\n\\includegraphics[scale=.5]{images\/Tangle1} \\,\\, \\raisebox{.6cm}{$\\cdot$} \\,\\,\n\\labellist\n\\small\n\\pinlabel $\\beta$ at 45 38\n\\endlabellist\n\\includegraphics[scale=.5]{images\/Tangle1}\n\\quad\\raisebox{.6cm}{$=$} \\quad\n\\labellist\n\\small\n\\pinlabel $\\beta$ at 45 38\n\\pinlabel $\\alpha$ at 120 38\n\\endlabellist\n\\includegraphics[scale=.5]{images\/Tangle2}\n\\]\n\n\\begin{Definition} Let $\\mathcal{R}$ be a coefficient ring containing $\\mathbb {Z}[z^{\\pm1}]$ as a subring. Define the {\\it Legendrian BMW algebra}, $\\operatorname{BMW}_n^\\mathfrak{Leg}$, as an $\\mathcal{R}$-module to be the quotient $\\mathcal{R} \\mathfrak{Leg}_n\/\\mathcal{S}$ where $\\mathcal{R} \\mathfrak{Leg}_n$ is the free $\\mathcal{R}$-module generated by Legendrian isotopy classes of Legendrian $n$-tangles in $J^1[0,1]$, and $\\mathcal{S}$ is the $\\mathcal{R}$-submodule generated by the ruling polynomial skein relations from Fig.~\\ref{fig:psuedo-gen}. Multiplication of $n$-tangles induces an $\\mathcal{R}$-bilinear product on $\\operatorname{BMW}_n^\\mathfrak{Leg}$.\n\\end{Definition}\n\nIn the remainder of the article, we fix the coefficient ring $\\mathcal{R}$ to be $\\mathbb {Z}\\big[s^{\\pm1}\\big]$ localized to include denominators of the form $s^n-s^{-n}$ for $n\\geq 1$ where $z = s- s^{-1}$. In Section~\\ref{sec:5}, we will work with the alternate variable $s= q^{1\/2}$.\n\n\\begin{figure} \\centering\n $\\sigma_i:=\\raisebox{-.75cm}{\\includegraphics[scale=.5]{images\/sigmai.png}}\\hspace{.75cm} \\hspace{.75cm} e_i:=\\raisebox{-.75cm}{\\includegraphics[scale=.5]{images\/ei.png}}$\n \\caption{Crossing and hook elements in $\\operatorname{BMW}_n^\\mathfrak{Leg}$.} \\label{fig:sigma}\n\\end{figure}\n\nFig.~\\ref{fig:sigma} indicates crossing and hook elements, $\\sigma_i, e_i \\in \\operatorname{BMW}_n^\\mathfrak{Leg}$, for $1 \\leq i < n$. Note that the fishtail and double-crossing relations imply\n\\begin{gather}\n \\sigma_i e_i = e_i \\sigma_i =0, \\qquad \\mbox{and} \\label{eq:fish2} \\\\\n \\sigma_i^2 = 1 + z \\sigma_i - z e_i, \\qquad \\mbox{for $1 \\leq i h(y) >h(z)$, and note that $z$ is the crossing from $\\beta$. The DGAs before and after the triple point move are related by a DGA isomorphism $\\phi\\colon (\\mathcal{A}, \\partial) \\rightarrow (\\mathcal{A}', \\partial')$ that maps~$x$ to an element of the form $x \\pm yz$ or $x\\pm zy$ and fixes all other generators. In particular, $\\phi$ restricts to the identity on the sub-algebra generated by the $Y$-generators.\nComposing all of the DGA isomorphisms from handleslide disks, we see that there is a DGA isomorphism $\\varphi\\colon \\big(\\mathcal{A}\\big(S^1_{xy}(K, \\beta)\\big), \\partial_1\\big) \\rightarrow \\big(\\mathcal{A}\\big(S^2_{xy}(K, \\beta)\\big), \\partial_2\\big)$ that restricts to the identity on all $Y$-generators. As a result, $\\varphi^*\\colon \\overline{{\\rm Aug}}_1\\big(S^{2}_{xy}(K, \\beta), \\mathbb{F}_q\\big) \\rightarrow \\overline{{\\rm Aug}}_1\\big(S^{1}_{xy}(K, \\beta), \\mathbb{F}_q\\big)$, $\\varphi^*\\epsilon = \\epsilon \\circ \\varphi$ induces the required bijection between $\\overline{{\\rm Aug}}_1\\big(S^{1}_{xy}(K, \\beta), \\mathbb{F}_q\\big)_{Y=0}$ and $\\overline{{\\rm Aug}}_1\\big(S^{2}_{xy}(K, \\beta), \\mathbb{F}_q\\big)_{Y=0}$.\n\n{\\bf Step 3.} Establish that\n \\begin{gather} \\label{eq:Step3}\n\\big| \\overline{{\\rm Aug}}_1\\big(S^{2}_{xy}(K, \\beta), \\mathbb{F}_q\\big)_{Y=0}\\big| = \\big(q^{n(n-1)\/2}\\big)^\\ell\\big| \\overline{{\\rm Aug}}_1\\big(S^{1}_{xz}(K, \\beta), \\mathbb{F}_q\\big)_{Y=0}\\big|.\n\\end{gather}\n\nThe generators of $S^1_{xz}(K,\\beta)$ are identified with a subset of the generators of $S^2_{xy}(K,\\beta)$, and this leads to an algebra inclusion $i\\colon \\mathcal{A}\\big(S^1_{xz}(K,\\beta)\\big) \\rightarrow \\mathcal{A}\\big(S^2_{xy}(K,\\beta)\\big)$. In fact, it is not hard to check that~$i$ is a DGA homomorphism; see \\cite[Proposition~4.23]{NRSSZ} for a detailed explanation in the case where $\\beta$ is the identity braid. The difference between the two DGAs is that $\\mathcal{A}\\big(S^2_{xy}(K,\\beta)\\big)$ has additional generators of the form $c^k_{i,j}$ and $x^k_{i,j}$ with $1 \\leq k \\leq \\ell$, $1 \\leq i} O(10) {\\rm TeV} $ has been suggested as\nan attractive solution to the Polonyi problem.\n\nIn reference~\\cite{PLB342-105}, however, it has been also pointed out\nthat the mass density of the lightest superparticle (LSP) produced by\nthe decay of the Polonyi field may overclose our universe if LSP is\nstable. As we will see below, the mass density of LSP increases as the\nreheating temperature $T_R$ due to the decay of the Polonyi field\ndecreases. Therefore, a lowerbound on $T_R$ is derived requiring that\nthe present mass density of LSP should not exceed the critical density\nof the universe $\\rho_c$. In this letter, we obtain the lowerbound on\n$T_R$ in the framework of the minimal SUSY SU(5) model with no-scale\ntype boundary conditions on the SUSY breaking parameters.\n\n\n\n\\section{The Model}\n\\label{sec:model}\n\nBefore starting cosmological arguments, let us first describe our\nbasic assumptions. We consider the minimal SUSY SU(5) model with\nno-scale type boundary conditions. This model has three types of Higgs\nfield; $H({\\bf 5})$ and $\\bar{H}({\\bf 5^*})$ which contain flavor\nHiggses $H_f$ and $\\bar{H}_f$, and $\\Sigma ({\\bf 24})$ whose\ncondensation breaks the SU(5) group into the gauge group of the\nminimal SUSY standard model (MSSM), $\\rm SU(3)_C\\times SU(2)_L\\times\nU(1)_Y$. For the Higgs sector, the superpotential is given by\n\\begin{equation}\n W = \\frac{1}{3} \\lambda {\\rm tr} \\Sigma^3 \n + \\frac{1}{2} M_\\Sigma {\\rm tr} \\Sigma^2\n + \\kappa \\bar{H} \\Sigma H\n + M_H \\bar{H} H,\n \\label{W_su5}\n\\end{equation}\nwhere $\\lambda$ and $\\kappa$ are dimensionless constants, while\n$M_\\Sigma$ and $M_H$ are mass parameters which are of the order of the\ngrand unified theory (GUT) scale $M_{\\rm GUT} (\\sim\n10^{16} {\\rm GeV} )$. Furthermore, the model also has the soft SUSY breaking\nterms;\n\\begin{equation}\n {\\cal L}_{\\rm soft} = \n - \\frac{1}{3} \\lambda A_\\Sigma {\\rm tr} \\Sigma^3 \n - \\frac{1}{2} M_\\Sigma B_\\Sigma {\\rm tr} \\Sigma^2\n - \\kappa A_H \\bar{H} \\Sigma H\n - M_H B_H \\bar{H} H +h.c.,\n \\label{L_soft_su5}\n\\end{equation}\nwhere $A_\\Sigma$, $B_\\Sigma$, $A_H$ and $B_H$ are SUSY breaking\nparameters. Minimising the Higgs potential, we find the following\nstationary point;\n\\begin{equation}\n \\vev{\\Sigma} = \n \\frac{1}{\\lambda} \\left \\{ \n M_\\Sigma + 2 \\left ( A_\\Sigma - B_\\Sigma \\right ) \n + O \\left ( \\frac{A_\\Sigma}{M_\\Sigma}, \\frac{B_\\Sigma}{M_\\Sigma} \\right ) \n \\right \\} \\times {\\rm diag}(2,2,2,-3,-3),\n \\label{vac_su5}\n\\end{equation}\nwhere the SU(5) is broken down to $\\rm SU(3)_C\\times SU(2)_L\\times\nU(1)_Y$. Regarding this stationary point as the vacuum, we obtain MSSM\nas the effective theory below the GUT scale $M_{\\rm GUT}$. Here, the\nmasslessness of the flavor Higgses $H_f$ and $\\bar{H}_f$ is achieved\nby a fine tuning among several parameters; $M_H - 3 \\kappa\nM_\\Sigma\/\\lambda \\simeq \\mu_H$, where $\\mu_H$ is the SUSY-invariant\nHiggs mass in MSSM.\n\nIn the present model, the parameters in MSSM at the electroweak scale\nis obtained by solving renormalization group equations (RGEs). Our\nmethod is as follows. The boundary conditions on the parameters in the\nminimal SUSY SU(5) model are given at the gravitational scale $M$.\nSince we assume the no-scale type supergravity models, all the SUSY\nbreaking parameters except for the gaugino mass vanish at the\ngravitational scale. From the gravitational scale to the GUT scale,\nthe parameters follow the renormalization group flow derived from RGEs\nin the minimal SUSY SU(5) model. Then we determine the parameters in\nMSSM at the GUT scale through an appropriate matching condition\nbetween the parameters in the SUSY SU(5) model and those in MSSM.\nFinally, we use RGEs in MSSM from the GUT scale to the electroweak\nscale in order to obtain the low energy parameters.\n\nAs for the matching condition, we have a comment. In the stationary\npoint (\\ref{vac_su5}), the mixing soft mass term of the two flavor\nHiggs bosons, $m_{12}^2\\bar{H}_fH_f$, is generated at the tree level,\nwhere $m_{12}^2$ is given by\n\\begin{equation}\n m_{12}^2 (M_{\\rm GUT}) \\simeq \n \\left [ \\frac{6\\kappa}{\\lambda} \n (A_\\Sigma - B_\\Sigma)(A_H - B_\\Sigma) \n -\\mu_H B_H.\n \\right ] _{\\mu =M_{\\rm GUT}}\n \\label{m12^2}\n\\end{equation}\nSince the mixing mass term depends on unknown parameters, $\\lambda$\nand $\\kappa$ in eq.(\\ref{W_su5}), we regard $m_{12}^2$ as a free\nparameter taking account of the uncertainty of $\\lambda$ and $\\kappa$\nin our analysis. Then, the low energy parameters are essentially\ndetermined by the gauge and Yukawa coupling constants and the\nfollowing three parameters; the supersymmetric Higgs mass $\\mu_H$, the\nmixing mass of the two flavor Higgs bosons $m_{12}^2$, and the unified\ngaugino mass.\\footnote\n{In fact, parameters in MSSM slightly depend on the parameters in the\nSUSY GUT such as $\\lambda$, $\\kappa$ and so on. In our numerical\ncalculation, we ignore the effects of these parameters on the \nrenormalization group flow.}\nHowever, it is more convenient to express these parameters by other\nphysical ones. In fact, one combination of them is constrained so that\nthe flavor Higgs bosons have correct vacuum expectation values;\n$\\langle H_f\\rangle^2+\\langle\\bar{H}_f\\rangle^2\\simeq (174 {\\rm GeV} )^2$. As\nthe other two physical parameters, we use the mass of LSP, $m_{\\rm\nLSP}$, and the vacuum angle\n$\\tan\\beta\\equiv\\langle{H_f}\\rangle\/\\langle{\\bar{H}_f}\\rangle$. Thus,\nonce we fix $m_{\\rm LSP}$ and $\\tan\\beta$, we can determine all the\nparameters in MSSM.\\footnote\n{Yukawa coupling constants are determined so that the fermions have\ncorrect masses. The gauge coupling constants are also fixed so that\ntheir correct values at the electroweak scale are reproduced.}\n\nFollowing the above procedure, we solve the RGEs numerically, and\nobtain the low energy parameters in MSSM. Then, we determine the mass\nspectrum and the mixing angles for all superparticles. One remarkable\nthing is that {\\it LSP almost consists of bino $\\tilde{B}$ which is\nthe superpartner of the gauge field for $U(1)_Y$} if we require that\nLSP is neutral. Therefore, in our model, the LSP mass $m_{\\rm LSP}$ is\nessentially equivalent to the bino mass. This fact simplifies the\nfollowing analysis very much.\n\n\\section{Density of LSP}\n\\label{sec:density}\n\nNow we are in a position to discuss the mass density of LSP produced\nby the decay of the Polonyi field. The decay of the Polonyi field\nproduces a large number of superparticles, which promptly decay into\nLSPs. The number density of LSP produced by the decay, $n_{{\\rm\nLSP},i}$, is of the same order of that of the Polonyi field\n$n_\\phi\\equiv\\rho_\\phi \/m_\\phi$ (with $\\rho_\\phi$ being the energy\ndensity of the Polonyi field). Just after the decay of the Polonyi\nfield, the yield variable for LSP, $Y_{\\rm LSP}$, which is defined by\nthe ratio of the number density of LSP to the entropy density $s$, is\ngiven by\n\\begin{eqnarray}\n m_{\\rm LSP} Y_{\\rm LSP} & \\simeq & \\frac{\\rho_\\phi}{s}\n \\simeq \\frac{m_{\\rm LSP}\\rho_{{\\rm LSP},i}}{m_\\phi s}\n \\sim \\left(\\frac{m_{\\rm LSP}T_R}{m_\\phi}\\right)\n \\nonumber \\\\\n & \\sim & 10^{-5} {\\rm GeV} \\left(\\frac{m_{\\rm LSP}}{100 {\\rm GeV} }\\right)\n \\left(\\frac{T_R}{1 {\\rm MeV} }\\right)\\left(\\frac{10 {\\rm TeV} }{m_\\phi}\\right),\n \\label{mY_LSP}\n\\end{eqnarray}\nwhere $\\rho_{{\\rm LSP},i}$ is the mass density of LSP just after the\ndecay of the Polonyi field. If LSP is stable and the pair annihilation\nof LSP is not effective, $Y_{\\rm LSP}$ is conserved until today.\nComparing the ratio given in eq.(\\ref{mY_LSP}) with the ratio of the\ncritical density $\\rho_c$ to the present entropy density $s_{0}$,\n\\begin{equation}\n \\frac{\\rho_c}{s_{0}} \\simeq 3.6 \\times 10^{-9}h^2~ {\\rm GeV} ,\n \\label{critical}\n\\end{equation}\nwhere $h$ is the Hubble constant in units of 100km\/sec\/Mpc, we see\nthat LSP overcloses the universe in the wide parameter region for\n$m_{\\rm LSP}, m_{\\phi}$ and $T_R$ which we are concerned with.\n\nIf the pair annihilation of LSP takes place effectively, its abundance\nis reduced to \n\\begin{equation}\n \\frac{n_{\\rm LSP}}{s} \\simeq \n \\left. \\frac{H}{s\\langle\\sigma_{\\rm ann}v_{\\rm rel}\\rangle} \n \\right | _{T=T_R},\n \\label{abundance_LSP}\n\\end{equation}\nwhere $\\sigma_{\\rm ann}$ is the annihilation cross section, $v_{\\rm\nrel}$ is the relative velocity, and $\\langle\\cdots\\rangle$ represents\nthe average over the phase space distribution of LSP. From\neq.(\\ref{abundance_LSP}), we obtain a lowerbound on the annihilation\ncross section,\n\\begin{equation}\n \\langle\\sigma_{\\rm ann}v_{\\rm rel}\\rangle \\mathop{}_{\\textstyle \\sim}^{\\textstyle >} \n 3\\times 10^{-8}h^{-2} {\\rm GeV} ^{-2}\n \\left( \\frac{m_{\\rm LSP}}{100 {\\rm GeV} }\\right)\n \\left( \\frac{100 {\\rm MeV} }{T_R}\\right),\n \\label{sv_limit}\n\\end{equation}\nin order that the mass density of LSP does not overclose the universe. \n\nComparing this bound with the annihilation cross section of LSP, we\nderive a bound on the reheating temperature by the decay of the\nPolonyi field. Since LSP is most dominated by bino, it annihilates\ninto fermion pairs. The annihilation cross section is given\nby~\\cite{PLB230-78}\n\\begin{equation}\n \\langle\\sigma_{\\rm ann}v_{\\rm rel}\\rangle \n = a + b\\langle v^2\\rangle,\n \\label{sigma*v}\n\\end{equation}\nwhere $\\langle v^2\\rangle$ is the average velocity of LSP,\nand\n\\begin{eqnarray}\n a & \\simeq &\n \\frac{32\\pi\\alpha_1^2}{27} \n \\frac{m_t^2}{(m_{\\tilde{t}_R}^2 + m_{\\rm LSP}^2 - m_t^2)^2}\n \\left ( 1 - \\frac{m_t^2}{m_{\\rm LSP}^2} \\right ) ^{1\/2} \n \\theta (m_{\\rm LSP}-m_t),\n \\label{s-wave} \\\\\n b &\\simeq& \\frac{8\\pi\\alpha_1^2}{3} \\sum_{m_f\\leq T} Y_f^4 \\left \\{ \n \\frac{m_{\\rm LSP}^2}{(m_{\\rm LSP}^2+m_{\\tilde{f}}^2)^2}\n - \\frac{2m_{\\rm LSP}^4}{(m_{\\rm LSP}^2+m_{\\tilde{f}}^2)^3}\n + \\frac{2m_{\\rm LSP}^6}{(m_{\\rm LSP}^2+m_{\\tilde{f}}^2)^4} \\right \\} .\n \\label{p-wave}\n\\end{eqnarray}\nHere, $\\alpha_1^2\\equiv g_1^2\/4\\pi\\simeq 0.01$ represents the fine\nstructure constant for U(1)$_{\\rm Y}$, $m_t$ the top-quark mass, $Y_f$\nthe hypercharge of the fermion $f$, and $m_{\\tilde{f}}$ the mass of\nthe sfermion $\\tilde{f}$. Notice that $a$ and $b$ terms correspond to\n$s$- and $p$-wave contributions, respectively. Taking\n$m_{\\tilde{f}}\\sim m_{\\rm LSP}\\sim 100 {\\rm GeV} $, the annihilation cross\nsection given in eq.(\\ref{sigma*v}) is at most $3\\times\n10^{-8} {\\rm GeV} ^{-2}$. Using this result in the inequality\n(\\ref{sv_limit}), we can see that the reheating temperature must be\nhigher than about 100MeV even if $\\langle v^2\\rangle\\sim 1$. If the\naverage velocity is smaller than 1, the constraint becomes more\nstringent, as we will see below.\n\n\n\n\\section{Thermalization of LSP}\n\\label{sec:thermalization}\n\n\nIn order to obtain the precise lowerbound on the reheating temperature\n$T_R$, we have to know $\\langle v^2\\rangle$, as well as the mass\nspectrum of the superparticles on which the annihilation cross section\ndepends. First, let us discuss the averaged velocity of LSP, $\\langle\nv^2\\rangle$. Since LSP is mostly the bino, it loses its energy by\nscattering off the background fermions. In the model with the no-scale\ntype boundary conditions, right-handed sleptons become the lightest\namong the sfermions, and hence LSP loses its energy mainly by\nscattering off the background electron (and $\\mu$ and $\\tau$, if the\ntemperature is higher than their masses). If LSP is relativistic, the\ncross section for this process, $\\sigma_{\\rm scatt}$, is estimated as\n\\begin{equation}\n \\langle \\sigma_{\\rm scatt} v_{\\rm rel} \\rangle \n \\simeq 128\\pi \\alpha_1^2\n \\frac{E_{\\rm LSP}^2 T_R^2}{m_{\\tilde{e}_R}^4 m_{\\rm LSP}^2},\n \\label{sigma_scatt}\n\\end{equation}\nwhere $E_{\\rm LSP}$ is the energy of LSP, and $m_{\\tilde{e}_R}$ the\nmass of the right-handed selectron.\\footnote%\n{This cross section is applied for $E_{\\rm LSP} T_{R} \\ll\nm_{\\tilde{e}_R}^2$.}\nThe energy loss rate for\nthe relativistic LSP, $\\Gamma_{\\rm scatt}^{\\rm R}$, is given by\n\\begin{equation}\n \\label{loss-rate}\n \\Gamma_{\\rm scatt}^{\\rm R} \\simeq \n n_e \\langle \\sigma_{\\rm scatt} v_{\\rm rel} \\rangle \n \\frac{\\Delta E_{\\rm LSP}}{E_{\\rm LSP}},\n\\end{equation}\nwhere $n_e$ represents the number density of the background electron\nand $\\Delta E_{\\rm LSP}$ is the averaged energy loss of LSP in \none scattering which is given by\n\\begin{equation}\n \\label{energy-loss}\n \\Delta E_{\\rm LSP} \\simeq \n -12E_{\\rm LSP}\\left(\\frac{T_R E_{\\rm LSP}}{m_{\\rm LSP}^2}\\right).\n\\end{equation}\nTaking the ratio of the energy loss rate $\\Gamma_{\\rm\nscatt}^{\\rm R}$ to the expansion rate $H$ of the universe, we find\n\\begin{equation}\n \\left. \\frac{\\Gamma_{\\rm scatt}^{\\rm R}}{H}\n \\right | _{E_{\\rm LSP}\\gg m_{\\rm LSP}}\n \\simeq \n 2\\times10^3 \\left( \\frac{E_{\\rm LSP}}{10^2 {\\rm GeV} } \\right)^3\n \\left( \\frac{T_R}{100 {\\rm MeV} } \\right)^4\n \\left( \\frac{100 {\\rm GeV} }{m_{\\tilde{e}_R}} \\right)^4\n \\left( \\frac{100 {\\rm GeV} }{m_{\\rm LSP}} \\right)^4.\n\\end{equation}\nThus, if $T_R \\mathop{}_{\\textstyle \\sim}^{\\textstyle >} $ a few $\\times$ 10MeV, the energetic LSP loses its\nenergy through the scattering off thermal electrons efficiently for\n$m_{\\tilde{e}_R}\\sim m_{\\rm LSP}\\simeq O(100\/GEV)$, and becomes a\nnon-relativistic particle.\n\n\nThe non-relativistic LSP further loses its energy by scattering off\nbackground electrons. The averaged loss of the kinetic energy for the\nnon-relativistic LSP in one scattering process, $\\Delta \\epsilon_{\\rm\nLSP}$, is given\nby\\footnote{%\nNaively, it is expected that $\\Delta \\epsilon_{\\rm LSP}$ is $\\sim\nT_R$. However this order of the energy loss is cancelled out when the\naverage is taken over angles of the incident particles and the actual\nenergy loss is much smaller than the naive expectation.}\n\\begin{equation}\n \\Delta \\epsilon_{\\rm LSP} \\simeq\n -\\frac{20\\epsilon_{\\rm LSP} T_R}{m_{\\rm LSP}}\n \\left( 1 -\\frac{T_R}{\\epsilon_{\\rm LSP}}\\right),\n \\label{Delta_E}\n\\end{equation}\nwhere $\\epsilon_{\\rm LSP}\\equiv E_{\\rm LSP}-m_{\\rm LSP}$ is the\nkinetic energy of LSP. As one can see in eq.(\\ref{Delta_E}), the LSP\nwhich has a kinetic energy larger than $\\sim T_R$ tends to lose its\nenergy through the scattering process, while LSP receives energy from\nthe thermal bath if its energy is smaller than $\\sim T_R$. Thus, if\nthe scattering processes take place effectively, the averaged kinetic\nenergy of LSP becomes $\\sim T_R$, {\\it i.e.} LSP goes into the\nkinetic equilibrium.\n\nThe energy loss rate $\\Gamma_{\\rm scatt}^{\\rm NR}$ for the \nnon-relativistic LSP is given by\n\\begin{equation}\n \\Gamma_{\\rm scatt}^{\\rm NR} \\simeq \n n_e \\langle\\sigma_{\\rm scatt} v_{\\rm rel}\\rangle \\times\n \\frac{20T_R}{m_{\\rm LSP}}\n \\simeq \n \\frac{5760\\alpha_1^2}{\\pi}\n \\frac{T_R^6}{m_{\\rm LSP}m_{\\tilde{e}_R}^4}.\n \\label{gamma_scatt}\n\\end{equation}\nThe LSP goes into the kinetic equilibrium if the scattering rate\n$\\Gamma_{\\rm scatt}^{\\rm NR}$ is larger than the expansion rate of the\nuniverse. Taking $m_{\\tilde{e}_R}\\sim m_{\\rm LSP}$, the ratio of\n$\\Gamma_{\\rm scatt}^{\\rm NR}$ to the expansion rate of the universe,\n$H$, is given by\n\\begin{equation}\n \\left. \\frac{\\Gamma_{\\rm scatt}^{\\rm NR}}{H} \\right | _{E_{\\rm LSP}\n \\sim m_{\\rm LSP}}\n \\simeq 4 \\times 10^{3} ~\n \\left ( \\frac{m_{\\rm LSP}}{100 {\\rm GeV} } \\right ) ^{-5}\n \\left ( \\frac{T_R}{100 {\\rm MeV} } \\right ) ^{4}.\n\\end{equation}\nThus, if the reheating temperature is higher than about 10MeV,\nproduced LSPs go into kinetic equilibrium as far as $m_{\\rm LSP} \\sim\nO(100)$GeV. Furthermore, as we discussed in the previous section,\nthe reheating temperature should be higher than at least 100MeV in\norder to decrease the mass density of LSP sufficiently. Thus, we\nconclude that the produced LSPs go into kinetic equilibrium if we\nrequire that the mass density of the relic LSP should not overclose\nthe universe.\\footnote\n{By the numerical calculation we have checked, in fact, that the\nscattering rate given in eq.(\\ref{gamma_scatt}) is always larger than\nthe expansion rate of the universe when the relic LSP does not\noverclose the universe. (See fig.2.)}\nIn this case, the averaged velocity is given by\n\\begin{equation}\n \\langle v^2\\rangle \\simeq \\frac{3T_R}{m_{\\rm LSP}}.\n\\end{equation}\n>From this we easily see that the LSP abundance given in\neq.(\\ref{abundance_LSP}) decreases as the reheating temperature gets\nhigher. Thus, we obtain the lowerbound on the reheating temperature.\n\n\n\\section{Results}\n\\label{sec:results}\n\nOnce we know the averaged velocity $\\langle v^2\\rangle$, we can\ncalculate the annihilation cross section of LSP, and get the\nlowerbound on the reheating temperature after the decay of the Polonyi\nfield. In this letter, we first solve RGEs based on the minimal SU(5)\nmodel with the no-scale boundary conditions, and determine the mass\nspectrum of the superparticles. We only investigate the parameter\nspace which is not excluded by the experimental or theoretical\nconstraints. The constraints which we use are as follows:\n\\begin{itemize}\n\\item Higgs bosons $H_f$ and $\\bar{H}_f$ have correct vacuum\nexpectation values.\n\\item Perturbative picture is valid below the gravitational scale.\n\\item LSP is neutral.\n\\item Sfermions (especially, charged sleptons) have masses larger than\nthe experimental lower limits~\\cite{PDG}.\n\\item The branching ratio for $Z$-boson decaying into neutralinos is \nnot too large~\\cite{PLB350-109}.\n\\end{itemize}\nThen, with the obtained mass spectrum of superparticles, we calculate\nthe annihilation cross section and determine the lowerbound on the\nreheating temperature from the following equation;\n\\begin{equation}\n \\left. \\frac{H}{s\\langle\\sigma_{\\rm ann}v_{\\rm rel}\\rangle} \n \\right | _{T=T_R} \\leq\n \\frac{\\rho_c}{s_0} \\simeq 3.6h^2 \\times 10^{-9} {\\rm GeV} .\n\\end{equation}\n\nIn fig.~1, we show the lowerbound on the reheating temperature in the\n$\\tan\\beta$ vs. $m_{\\rm LSP}$ plane. In the figures, large or small\n$\\tan\\beta$'s are not allowed since the Yukawa coupling constant for\nthe top quark or bottom quark blows up below the gravitational scale\nfor such $\\tan\\beta$'s. Furthermore, there also exists a lowerbound on\nthe LSP mass. In the case where $\\tan\\beta \\mathop{}_{\\textstyle \\sim}^{\\textstyle <} 20$, charged sfermions\nbecome lighter than the experimental limit if the LSP mass becomes\nlighter than $\\sim 50 {\\rm GeV} $. On the other hand, for the large\n$\\tan\\beta$ case, unless the bino mass is sufficiently large, the\nlightest charged slepton becomes LSP. (Remember that the dominant\ncomponent of LSP is bino.) Thus, the lowerbound on $m_{\\rm LSP}$ is\nobtained. As we can see, the reheating temperature should be larger\nthan about 100MeV, even for the case where $m_{\\rm LSP}\\sim 50 {\\rm GeV} $.\nThe constraint becomes more stringent as $m_{\\rm LSP}$ increases,\nsince the masses of the superparticles which mediate the annihilation\nof LSP becomes larger as the LSP mass increases.\n\nIf we translate the lowerbound on the reheating temperature into that\nof the Polonyi mass $m_\\phi$, we obtain $m_\\phi \\mathop{}_{\\textstyle \\sim}^{\\textstyle >} 100 {\\rm TeV} $ (see\neq.(\\ref{rtemp})). We can also see that the lowerbound is almost\nindependent of $\\tan\\beta$. In fig.~2, We show the lowerbound on $T_R$\nas a function of the LSP mass for $\\tan\\beta =10$, and $\\mu_H >0$.\n\nHere, we should comment on the accidental case where the annihilation\nprocess hits the Higgs pole in the $s$-channel. If the LSP mass is\njust half of the lightest Higgs boson mass, the LSP annihilation cross\nsection is enhanced since LSP has small but nonvanishing fraction of\nhiggsino component. If the parameters are well tuned, such a situation\ncan be realized and the lowerbound of $T_R$ decreases to $O(10) {\\rm MeV} $.\nHowever, we consider that such a scenario are very unnatural since a\nprecise adjustment of the\nparameters is required in order to hit the Higgs pole.\\footnote\n{In the case where the annihilation process hits the pole of heavier\nHiggs bosons, the cross section is not enhanced so much, since the\nwidths of the heavier Higgs bosons are quite large.}\n\n\\section{Conclusions}\n\\label{sec:conclusions}\n\n\nIn this letter, we have obtained the lowerbound on the reheating\ntemperature due to the decay of the Polonyi field in a framework of\nthe no-scale type supergravity model. As a result, we have seen that\nthe Polonyi mass should be larger than about 100TeV which may raise a\nnew fine-tuning problem~\\cite{PLB173-303}.\n\nWe have assumed the minimal SUSY GUT model in the present\nanalysis. However, the main conclusion is not changed as far as LSP\nis mostly the bino, because the minimum value of the lowerbound ($T_R\n\\simeq 100$MeV) is obtained when the mass of the selectron takes the\nexperimentally allowed lower limit.\n\nWe have assumed that LSP is stable so far. However, if we introduce\n$R$-parity violation, LSP becomes unstable and the allowed $T_R$ is as\nlow as a few MeV. It is also the case if we assume a very light LSP\nsuch as a neutral higgsino~\\cite{PLB131-59} or an\naxino~\\cite{NPB358-447} whose masses are less than about 100MeV.\n\n\\section*{Acknowledgement}\n\nTwo of the authors (T.M. and T.Y.) would like to thank M.~Yamaguchi for\nuseful discussions in the early stage of this work.\n\n\\newpage\n\\newcommand{\\Journal}[4]{{\\sl #1} {\\bf #2} {(#3)} {#4}}\n\\newcommand{Ap. J.}{Ap. J.}\n\\newcommand{Can. J. Phys.}{Can. J. Phys.}\n\\newcommand{Nuovo Cimento}{Nuovo Cimento}\n\\newcommand{Nucl. Phys.}{Nucl. Phys.}\n\\newcommand{Phys. Lett.}{Phys. Lett.}\n\\newcommand{Phys. Rev.}{Phys. Rev.}\n\\newcommand{Phys. Rep.}{Phys. Rep.}\n\\newcommand{Phys. Rev. Lett.}{Phys. Rev. Lett.}\n\\newcommand{Prog. Theor. Phys.}{Prog. Theor. Phys.}\n\\newcommand{Sov. J. Nucl. Phys.}{Sov. J. Nucl. Phys.}\n\\newcommand{Z. Phys.}{Z. Phys.}\n","meta":{"redpajama_set_name":"RedPajamaArXiv"}} +{"text":"\\section{Introduction\\label{ch5:int}}\n\nThis work is aimed at the question of what provides the\ndominant power source of ultraluminous infrared galaxies, that\nis galaxies with luminosities in the wavelength range\n8--1000 \\hbox{\\micron~} in excess of $10^{12}$ \\hbox{L$_\\odot$}~($H_0$=75 km\ns$^{-1}$ Mpc$^{-1}$). This question is of interest because\nit is possible that the power source of these galaxies is\nqualitatively different from the dominant power source of most\ninfrared galaxies of lower luminosity. The observational\napproach of this work is the use of moderate resolution\n8--13 \\hbox{\\micron~} spectroscopy. \n\nIn very broad strokes, the context of the debate over the\nquestion addressed here can be summarized as \nan attempt, on the one hand, to explain ultraluminous infrared\ngalaxies as more powerful versions of starburst galaxies,\nthat is as super starbursts (Joseph \\&\nWright~1985), where massive star\nformation provides the power source, and on the other hand,\nto demonstrate that ultraluminous infrared galaxies are\nthe precursors of quasars (Sanders et al. 1988) and are\npowered by active galactic nuclei (AGNs).\n\nThe question has remained open to serious debate for\nessentially one reason: dust. Infrared galaxies by their\nnature radiate the bulk of their energy through the \ndegradation of the optical and UV photons emitted by the\nprimary power source into infrared photons emitted by dust\nheated to a temperature of $\\sim$50 K. This has the\neffect of erasing most of the commonly accepted\nfeatures used to distinguish the characteristics of\nstarbursts and AGNs.\n\nOptical and, to a lesser degree, near-infrared studies are\nparticularly susceptible to this difficulty since spectroscopic signatures of AGNs in ultraluminous\ninfrared galaxies should only betray sources that are\nrelatively unobscured and therefore might not contribute\nsubstantially to the heating of the dust that is the final source\nof most of the observed emission. However, Veilleux et\nal. (1995) have shown in a large sample of infrared galaxies\nthat the fraction of infrared galaxies with AGN-like optical\nspectra increases as a function of increasing luminosity, providing\nimportant circumstantial evidence that AGN-like activity is\nassociated with ultraluminous infrared galaxies. The strength\nof this evidence might be considered to be weakened, however, by\nthe finding of Goldader et al. (1995) that the ratio of\nBr$\\gamma$ to infrared luminosity decreases as a function of\nluminosity, suggesting that dust obscuration plays an even\ngreater role in ultraluminous infrared galaxies than in\nluminous infrared galaxies. This suggests the perverse\npossibility that energetically dominant starbursts in\nultraluminous infrared galaxies are so obscured in the optical\nthat only incidental unobscured AGNs are detected.\n\nRadio observations are unaffected by dust obscuration and can\nprovide high spatial resolution. However, the primarily\nsynchrotron radiation that is detected is only indirectly linked\nto the power source of infrared galaxies. In starbursts, it is\nthought that the radio emission arises from electrons\naccelerated in supernovae remnants, while in AGN, relativistic\njets are strong radio emitters. Condon et al. (1991; hereafter\nCHYT) have employed the high spatial resolution provided by the\nVery Large Array (VLA) to show that the radio sizes of a large\nfraction of ultraluminous infrared galaxies are comparable to\nthe minimum size required for a blackbody that reproduces the\n25, 60, and 100 \\hbox{\\micron~} flux densities of these sources. From this\nthey infer that starbursts rather than AGNs may power these\nsources. At even higher spatial resolution provided by very\nlong baseline interferometry (VLBI), Lonsdale, Smith \\& Lonsdale\n(1993) find evidence for radio structure on the scale of\nmilliarcseconds, some components of which have brightness\ntemperatures well in excess of what could be expected in\nstarburst-related emission. They deduce from this that AGNs are\npresent in five of eight of the ultraluminous infrared galaxies\nin their sample\\footnote{Smith, Lonsdale \\& Lonsdale (1998)\nsuggest that radio supernova remnants may be responsible for\nsome of the structure they detect}.\n\nIn this work, the dust emission itself is examined for\nspectral clues that can reveal the nature of the power\nsource in infrared galaxies. In a sample of 60 galaxies,\nRoche et al. (1991; hereafter RASW) have shown that their\n8--13 \\hbox{\\micron~} spectra can be classified into three types:\nthose with prominent polycyclic aromatic hydrocarbon (PAH)\nfeatures generally show \\ion{H}{2} region-like optical\nspectra, those with flat featureless spectra have Seyfert\n1 or quasar-like optical spectra, and those rare galaxies\nwith silicate absorption features usually have Seyfert 2 optical\nspectra. Dudley \\& Wynn-Williams (1997) have extended\nthe work on galaxies with very deep silicate features to\nshow that the size of the power source can be constrained\nto be smaller than a few parsecs, implying that such sources\nare powered by deeply embedded AGNs. Here, the 8--13 \\hbox{\\micron~}\nspectra of a well-defined sample of 25 luminous and\nultraluminous infrared galaxies are examined in this\ncontext. \n\nIt should be noted that all of the sources for which new\nspectral observations are reported here can be found in one or\nmore of four {\\em IRAS}-based catalogs: the Bright Galaxy Survey\n(Soifer et al. 1989), the Extended Bright Galaxy Survey\n(Sanders et al. 1995), the Two Jansky Survey (Strauss et\nal. 1992), or the Extended 12 \\hbox{\\micron~} Survey (Rush, Makan \\&\nSpinogio 1993).\n\nObservations and data reduction procedures are described in\nSection 2. Section 3 presents the new spectral observations. The\nclassification of the spectra is discussed in\nSection 4.1. The justification for the classification of the ultraluminous galaxies that fall within the later subsample\nis discussed in some detail in Section 4.2. In Section 4.3 a subsample\nof galaxies is defined and an apparent trend with luminosity\nwithin it is examined. The possibility of an\nanti-correlation between the equivalent width of the 11.3\n\\hbox{\\micron~} PAH feature and the ratio of infrared light to molecular\ngas mass is discussed in Section 4.4, and conclusions are\ndrawn in Section 5.\n\n\n\\section{Observations and Data Reduction\\protect\\label{ch5:odr}}\n\n\n\nThe new observations reported here were obtained over the\ncourse of 5 years, from 1991 through 1995, using the 10 and\n20 \\hbox{\\micron~} Cooled Grating Spectrograph (CGS3) mounted at the\nCassegrain focus of the United Kingdom Infrared Telescope\n(UKIRT) on Mauna Kea in Hawaii. Observations in 1991 were\nobtained by G. Wynn-Williams and J. Goldader, in 1992 by\nG. Wynn-Williams and C. Dudley, and in 1993--1995 by\nC. Dudley. For parts of each observing run a UKIRT\nscientist (J. Davies, T. Geballe, or G. Wright) was present\nat the summit and participated in the observations. CGS3 has\nbeen described by Cohen \\& Davies (1995). All but one of\nthe spectra reported here were obtained using the low-resolution spectroscopy mode with $\\lambda\/\\Delta\\lambda\n\\sim 60$ in a series of beam-switched, chopped observations\nwith a chop frequency of $\\sim$10 Hz and a beam switch\nfrequency of $\\sim$0.5 Hz. To fully sample each\nspectrum, two interlacing grating positions were observed\nwith changes between the two positions occurring after\n10--16 beam-switched pairs. Table 1 gives a log of\nobservations for the galaxies. In Table 1, column 1 is\nthe object name, columns 2 and 3 give the aperture centre\nin RA and DEC, column 4 gives the position reference,\ncolumns 5 and 6 give the object redshift and reference,\ncolumn 7 gives the date of observation, column 8 gives the\ncircular aperture size used in arcseconds (full width at\n10 per cent power), column 9 gives the chop throw in arcseconds,\nand column 10 gives the chop direction.\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\nObservations were calibrated in wavelength by means of\nhigh order spectral measurements of a Kr lamp in the\nmanner discussed by Hanner, Brooke \\& Tokunaga (1995).\nA master calibration spectrum was first selected and\ncalibrated in wavelength by flattening a 2$\\times$\nover-sampled Kr lamp spectrum (4 grating positions) with\nthe spectrum of a soldering iron replacing the Kr lamp in\nthe beam. Kr lamp spectra from each night's observations\nwere then cross correlated against the unflattened master\nspectrum to give a wavelength calibration for each night.\nTypical shifts from the nominal CGS3 wavelength\ncalibration were less than a resolution element, and were\nnever more than 2 resolution elements.\n\nCorrections for instrumental response, atmospheric\ntransmission, and flux calibration were performed by dividing\nsource spectra by the spectrum of a standard star with a\nknown 10 \\hbox{\\micron~} flux density and an assumed blackbody\ntemperature based on its spectral type, and then multiplying\nby the Planck curve of that temperature scaled to match the\nstandard star in flux density. Observations of standard\nstars were performed with the same aperture, chop throw,\nposition angle and frequency as the galaxy observations they\ncalibrate. The standard star observations were made to match\nin airmass so that no additional airmass corrections would be\nrequired. For the standard stars, the number of beam-switched pairs between grating shifts was typically 3--5. The\nstars have been chosen, where possible, to have a spectral\ntype of K0 or earlier to avoid the effects of SiO absorption,\nH$^-$ opacity, and circumstellar dust emission that can\neffect the spectral shape of late-type stars. Table 2 gives\nthe HR number, 10.1 \\hbox{\\micron~} flux density, photometric reference,\nand assumed blackbody temperature of each standard star, in\ncolumns 4, 5, 6, and 7, respectively, for each of the\nobservations given in Table 1. In column 2 of Table 2,\nalternative names are given for galaxies that are found in\nmultiple catalogs, while column 3 gives the date of each\nobservation as in Table 1. It should be noted that an error of $\\sim$1000 K in the assumed blackbody temperature of the standard\nstars leads to a $\\sim$2 per cent error in flux calibration at\neither end of the passband 8--13 \\hbox{\\micron~} which is negligible for\nthe purposes of this work, while for those stars of late\nspectral type, deviations from the assumed blackbody can be\n$\\sim$10 per cent in the SiO band (Cohen et al. 1995). Systematic\nuncertainties in overall flux calibration of 15--20 per cent can\narise due to telescope pointing errors with the stronger\neffects for the smaller apertures. For wavelengths shortward\nof 8.2 \\hbox{$\\mu$m}, strong systematic effects can occur as a\nresult of a variable water vapor column in the Earth's\natmosphere.\n\n\n\n\n\nAll spectra were regridded to a common wavelength scale using a\nGaussian filter of $\\sigma$ = 0.064 \\hbox{\\micron~}, which lead to a slight\ncorrelation between neighboring data points. Those spectra of the\nsame source obtained on two or more nights were combined,\nweighted by integration time. Data not falling in the ranges\n7.7--9.1 and 10.0--13.4 \\hbox{\\micron~} have been excluded due to poor\natmospheric transmission.\n\n\nSpectra of Wynn-Williams et al.'s (1991) Arp 299B1 were obtained on 1994 February 6 and 7\nusing CGS3 in high-resolution mode ($\\lambda\/\\Delta\\lambda\n\\sim 250$) using a 3\\farcs26 full width at 10 per cent power\ncircular aperture. The spectra obtained on each night were\nthe result of two sets of observations each consisting of two\ninterlacing grating positions but with different wavelength\nranges to cover the spectral range 10--13.2 \\hbox{$\\mu$m}. The chop\nthrow was 20\\arcsec, and the chop and beam switch PA were\n45$^\\circ$. About 30 per cent of the first night's observations were\naffected by clouds; these data were rejected by inspecting the\nchange in the dispersion of the total signal. The Kr lamp was\nobserved on each night to provide wavelength calibration. The\nstrongest lines from each night agreed in wavelength to 3 per cent of\na resolution element so that the Kr spectra of the two nights\nwere combined, and a difference in wavelength of 30 per cent of a\nresolution element was measured between the expected (based on\nthe nominal calibration of CGS3) and the measured wavelengths\nof 5 Kr lines. An additive correction of 0.02 \\hbox{\\micron~} was\napplied to the observations, this being the interpolation of\nthe 2.117 and 2.191 \\hbox{\\micron~} Kr lines measured at 6th order at the\nexpected position of the galaxy [\\ion{Ne}{2}] line. Flux\ncalibration was performed using HR 4067 with a 10.1 \\hbox{\\micron~} flux\ndensity of 103 Jy and an assumed temperature of 4000 K on the\nfirst night, and HR 5340 using a 10.1 \\hbox{\\micron~} flux density of 738\nJy and an assumed temperature of 4400 K on the second night.\nNo significant difference was found between the two nights'\nobservations, so they were combined according to their\nstatistical weights. The difference in airmass between the\nsource and standard observations was $\\sim$0.5 but monitoring\nof BS 4518 on the first night at airmasses similar to the\nsource observations showed that the level of spectroscopic\nvariation over the observed wavelength range as a function of\nairmass was negligible compared to the final signal-to-noise\nratio. The overall photometric calibration is estimated to be\ngood to about 25 per cent while the typical signal-to-noise ratio of\nan independent data point is about 10.\n\n\n\n\n\\section{Results\\label{ch5:r}}\n\n\n\n\n\nThe intensity and equivalent width of the 11.3 \\hbox{\\micron~} PAH\nfeature have been measured directly for all of the sources by\nestimating the continuum in the rest wavelength (see Table 1) ranges\n10.5--10.9 and 11.7--12.1 \\hbox{\\micron~} and subtracting the linear\ninterpolation of these two estimates from the spectra to\nmeasure the residual flux in the range 10.9--11.7 \\hbox{$\\mu$m}.\nUncertainties in the feature intensities represent the\ncombined uncertainties in the residual flux in this range but\nnot the uncertainties in continuum subtraction. For the\nequivalent width estimates, uncertainties in the continuum\nlevel have been included. In most cases the continuum\nuncertainties are dominant in the equivalent width estimates,\nleading to indeterminate results in the case of limits, but in\na few cases the uncertainty in the residual is the dominant\ncomponent, and upper limits to these equivalent widths can be\nset.\n\nSince model fitting can lead to substantially cleaner\nestimates of feature fluxes (of order the signal-to-noise\nratio in the peak spectral element being fit)\nestimates based on the model that is applied to these \ndata to test for the presence of a silicate absorption\nfeature are also given. The model consists of a\npower-law continuum and an additive PAH spectrum\nrepresented by the spectrum of the Orion Bar taken from\nRoche (1989). The fitting of this model was accomplished\nusing a version of Bevington's (1969) program CURVEFIT\nsupplied with IDL ver. 4.0 using three free parameters,\nthe power-law continuum scaling and index ($f_\\nu =\n\\beta\\times\\nu^{-\\alpha}$), and the scaling of the Orion Bar\nspectrum. The data were fit in the rest wavelength ranges\n10.0--10.9, 11.0--11.4, and 11.7--12.6 \\hbox{\\micron~} using initial\nguesses based on linear fits to the logarithm of data\nbinned in the ranges 10.0--10.5, 10.5--10.9, and\n11.7--12.6 \\hbox{\\micron~} for the power-law scaling and index, and\nbased on the continuum subtracted estimates of the flux in\nthe 11.3 \\hbox{\\micron~} feature. The weighting of the data was\nstatistical. As discussed recently by Goldader et\nal. (1995), the use of a power-law as a fitting function\ntends to throw the bulk of the errors in the fit\nparameters into the power-law scaling, so that the formal\nerrors of the remaining parameters are underestimated.\nFurther, the model assumptions introduce further\nsystematic uncertainties. In particular, by fitting the\ndata in the range 11.2--12.6 \\hbox{$\\mu$m}, it is assumed that\nthe ratio of 11.3 \\hbox{\\micron~} feature emission with respect to the\nunderlying plateau emission that extends to 12.9 \\hbox{\\micron~} is\nthe same as that found at the position of peak of the 11.3\n\\hbox{\\micron~} emission feature in the Orion Bar. This ratio,\nhowever, is know to vary spatially within the Orion Bar (Roche\net al. 1989), and also within at least one external galaxy\n(Dudley \\& Wynn-Williams 1999). The consequences\nof this uncertainty are largest in spectra where the PAH\nemission is most powerful relative to the continuum. Given\nthat the galaxy redshifts make it impossible to estimate the\ncontinuum beyond 13.1 \\hbox{\\micron~} (rest frame) in $\\sim$50 per cent of the\nsources, it has not been possible to quantify the importance of\nthis effect in these data, regardless of obvious\nsignal-to-noise issues. Due to these multiple and tangled\nuncertainties, both results of direct measurements and model\nfits with formal errors are presented, but it must be\ncautioned that the model parameter errors are seriously\nunderestimated. In particular, model 3 $\\sigma$ upper limits\nto the 11.3 \\hbox{\\micron~} intensity are roughly the same as measured 1\n$\\sigma$ upper limits to the 11.3 \\hbox{\\micron~} intensity. The new\nspectra and model fits are presented in Fig. 1.\n\n\n\n\n\n\n\n\n\nTable 3 presents the model power-law index in column 2,\nthe measured (upper) and model (lower) 11.3 \\hbox{\\micron~} feature\nintensity in column 3, the 11.3 \\hbox{\\micron~} feature equivalent width in\ncolumn 4, the ratio of the flux density of the source in\nthe rest wavelength range 8.0--8.4 \\hbox{\\micron~} to the model flux\ndensity in the same range where the errors represent the\nerrors in the estimates of the source flux densities only.\nMeasurements of fine structure lines are given in columns\n6 and 7. For the measurements of fine structure lines the\ncontinuum was estimated by eye, and for detections of the\n12.8 \\hbox{\\micron~} [\\ion{Ne}{2}] line it must be cautioned that a\nsignificant amount of the intensity could be due to a\nblended PAH emission feature at 12.7 \\hbox{\\micron~} except in the\ncase of Arp 299B1, where the spectral resolution is\nsufficient to distinguish the two features. Errors for\ndetections and limits are 1 $\\sigma$ and 3 $\\sigma$,\nrespectively, except in the case of the model 11.3 \\hbox{\\micron~}\nequivalent width, which is reported in all cases to give a\nsense of the model continuum uncertainties. In the case\nof Arp 299B1, the fine structure [\\ion{S}{4}] and\n[\\ion{Ne}{2}] results are derived from the high-resolution\nspectrum. The redshift of the [\\ion{Ne}{2}] line (2900\n\\hbox{km s$^{-1}$}) agrees within 1.4 $\\sigma$ (1 $\\sigma$ = 170 \\hbox{km s$^{-1}$}\nbased on Gaussian fitting) with the CO redshift measured\nby Sargent \\& Scoville (1991), while much more precise\nagreement between unpublished IRTF CSHELL \\hbox{Br$\\gamma$}~ data and the CO\nobservations is known, so that some\nconfidence in [\\ion{Ne}{2}] identification is warranted. It\nshould be noted\nthat [\\ion{Ne}{2}] is not detected in the low resolution\n8--13 \\hbox{\\micron~} spectrum of Arp 299B1 but that the 3 $\\sigma$\nupper limit for this feature is 2.1 $\\times 10^{-15} $ W\nm$^{-2}$, which is consistent with the intensity measured\nat high resolution. Further, the ratio of I(11.3) to\nI([\\ion{Ne}{2}]) (1.7) is not unusual for galaxies for\nwhich the ratio has previously been measured (RASW).\n\n\n\n\n\n\n\n\\addtocounter{table}{1}\n\n\n\n\n\n\n\n\nAs is apparent in Table 3, while a wide range of power-law\nindices are necessary to describe the long-wavelength\ncontinuum, the ratio of 8.2 \\hbox{\\micron~} flux to model 8.2 \\hbox{\\micron~}\nflux is not significantly different from unity for a large\nfraction of the galaxies even in cases where the power-law\nindex is large and a minimum near 9.7 \\hbox{\\micron~} is pronounced.\nAs a result, it is not generally necessary to invoke silicate\nabsorption to explain the observed spectra. \n\n\n\n\\section{Discussion\\label{ch5:d}}\n\n\\subsection{Spectral classification\\label{ch5:d1}}\n\nRASW have found that a three-type\nclassification scheme is sufficient to classify the 8--13\n\\hbox{\\micron~} spectra of most galaxies. Their scheme consists of\nflat-featureless, silicate absorption or emission, and PAH-type (their UIR)\nspectra. Further, there is a general correspondence\nbetween the 8--13 \\hbox{\\micron~} spectrum of a galaxy and its optical\nspectral classification based on optical emission lines. The\ncorrespondence is that those galaxies with flat 8--13 \\hbox{\\micron~}\nspectra are often broad-line AGNs, those with silicate\nabsorption are frequently narrow-line AGNs, and those with PAH\nfeatures typically have optical spectra with line ratios\nsimilar to Galactic\n\\ion{H}{2} regions. \n\nParenthetically, the identification of the mid-infrared\nemission feature analysed here is not entirely secure. Other\nlaboratory analogs, including quenched carbonaceous composites\n(Sakata \\& Wada~1989), hydrogenated amorphous\ncarbon (Duley~1989), and anthracite (Ellis et\nal.~1994), have been proposed in addition to\nthe PAH model (Puget \\& L\\'eger~1989) that is\nadopted here for ease of expression. However, the precise\nidentification of the chemical source of these features is not\nparticularly important for the present work. What is\nimportant is that the strength of one feature may be predicted\nto some extent from observations of another. In particular\nthe strength of the 7.7 \\hbox{\\micron~} feature may be predicted from\nobservations of the 11.3 \\hbox{\\micron~} feature to within about a factor of 2\n(Zavagno, Cox \\& Baluteau 1992).\n\nThe model fitting performed at the long-wavelength end of the\nspectra can be used effectively to test for the presence of\nstrong silicate absorption. Two cases might occur. First, a\nPAH spectrum might undergo external extinction. In this case,\ntwo effects will contribute to the model under predicting the\nobserved flux in the shorter wavelength half of a spectrum. The\nfirst effect is that the 11.3 \\hbox{\\micron~} feature will undergo more\nsevere extinction than the region near 8 \\hbox{\\micron~} due to the\nspectral shape the silicate component of the interstellar\nextinction curve. This has the effect of reducing the predicted\nPAH strength. The second effect is that the slope of the model\ncontinuum will have a larger spectral index than would otherwise\nbe the case if no silicate absorption were present. The model\nwould therefore under-predict the continuum underlying the 7.7\nand 8.6 \\hbox{\\micron~} feature complex. Both of these effects will cause\nthe model to under-predict the shorter wavelength observations.\nSecond, a PAH spectrum may undergo no particular extinction,\nhowever, a second silicate-absorbed continuum component may also\ncontribute to the spectrum. In this case, the model should\npredict the PAH component well. However, if the silicate\nabsorption is strong, the model spectral index will be increased\ndue to the additional component, and the continuum in the short\nwavelength half will be underestimated so that the model will\nalso under-predict the observed flux. \n\nThe RASW classification scheme is applied to the present data\nin the following way: the ratio of the observed to model flux\ndensity between 8.0 and 8.4 is used as a guide to classification\nsuch that if this ratio is 3 $\\sigma$ above 2.5, then the\nsource is classified as silicate. This cut may somewhat\nunderestimate the number of galaxies in which silicate\nabsorption is present compared to the approach taken in\nDudley \\& Wynn-Williams (1999), where the cut is taken at\n2.0. For the purposes of definite classification, under the\nworking hypothesis that all infrared galaxies are PAH-type\nunless convincingly shown to be otherwise, this more\nconservative value is appropriate. In cases where the 11.3\n\\hbox{\\micron~} feature is not formally detected but only allowed (dot\ndashed curves in Fig. 1), and where a minimum is present\nnear 9.7 \\hbox{$\\mu$m}, a question mark is appended to the PAH\ndesignation; however, whether or not there is a formal\ndetection of the 11.3 \\hbox{\\micron~} feature in sources with obviously\nflat spectra, the classification is given as flat. A mixed\nclassification of PAH and silicate is given where PAH\nemission can explain $\\sim$30 per cent of the observed emission\nbetween 8.0 and 8.4 \\hbox{$\\mu$m}. Under this condition, the\nclassification of Arp 220 by RASW as both PAH and silicate\nwould be revised to just silicate. Arp 220 is discussed\nin detail in Section 4.2.5.\n\n\n\n\n\n\n\n\n\\addtocounter{table}{1}\n\n\nTable 4 gives, for each source in column 1, the spectral\nclassification based on the present data and on optical\nobservations, along with 12--100 \\hbox{\\micron~} photometry, the infrared\nluminosity, and the star formation efficiency. In some cases\nthe optical spectral classification has been inferred from the\ndata of Veilleux et al. (1995), if some diagnostics were\nmissing. It is clear from Table 4 that the general\ncorrespondence between optical spectroscopy and 8--13 \\hbox{\\micron~}\nspectroscopy is quite similar to that found by RASW. Neglecting\nLINER-type galaxies, four exceptions where AGNs are identified\noptically but which are found to have PAH-type spectra in this\nwork are Mkn 1034, IRAS 04154+1755, IRAS 05189$-$2524, and Zw\n475.056.\n\nMkn 1034 is a known Seyfert 1 galaxy; however its 8--13 \\hbox{\\micron~}\nspectrum can be fit with a PAH plus continuum model. This may\nbe consistent with the finding of Sopp \\& Alexander (1992) that\nonly half of the 6-cm flux density measured in a 14\\farcs5 beam\nis also contained within a 1\\farcs5 beam, suggesting that\nspatially extended star formation plays a role in this source.\nA similar explanation may be invoked for the presence of PAH\nemission in the spectra of the Seyfert 2 galaxies IRAS\n04154+1755 and Zw 475.056. Crawford et al. (1996) find IRAS\n04154+1755 to be resolved at 6 cm using both 4-arcsec and\n0.5-arcsec beams. In the map of Zw 475.056 presented by CHYT,\nlow-level radio emission extends over 1.5-arcsec. IRAS\n05189$-$2524 is discussed in detail in the next section; however\nit should be noted here that the presence of exceptions, in this\nsample, to the the correspondence between optical and 8--13 \\hbox{\\micron~}\nspectral classification found by RASW should not be taken as\nevidence that the correspondence is merely coincidental. The\nsources presented by RASW were sometimes selected on the basis\nof optical spectroscopy to examine the question of what kind of\nspectrum they would present at 8--13 \\hbox{$\\mu$m}. Hence, many of\ntheir sources were archetypical examples of their optical class.\nIn the present sample, selection is based solely on infrared\nproperties, so that mixed cases might well be expected.\n\nIt is probably not coincidental that these four sources are\namong the sources in Table 3. with the smaller values of\n$\\alpha$ suggesting that the AGN in these sources may contribute\nsomewhat to the 8--13 \\hbox{\\micron~} continuum; strong 12 and 25 \\hbox{\\micron~}\nemission is known to select for AGNs that have unambiguous\noptical spectra (de Grijp et al. 1985; Spinoglio \\& Malkan, see also\nthe AGN spectral energy distributions presented by Sanders et\nal. 1989).\n\n\\subsection{Comments on individual ultraluminous infrared galaxies}\n\n\n\n\nThe observed properties of the ultraluminous infrared galaxies\npresented here show a surprising range. A closer look at these\nproperties will be discussed in this section. \n\n\\subsubsection{A phenomenological description of IRAS 05189$-$2524}\n\nOne of the greatest surprises in these data is the case of\nIRAS 05189$-$2524, which might have been expected to show a flat\nor silicate absorbed 8--13 micron spectrum based on its\noptical AGN-like spectrum (Sanders et al. 1988). What is\nobserved however is a spectrum that is well fit by a PAH plus\npower-law continuum model (see Fig. 1). Additionally, the\nstrength of the 11.3 \\hbox{\\micron~} emission relative to the 8--1000\n\\hbox{\\micron~} emission is quite similar to known starburst galaxies so that\nit seems possible that star formation might be the primary\npower source of the galaxy. Here, a tentative decomposition\nof its spectral energy distribution is proposed. In Fig.\n2, a comparison is made between IRAS 05189$-$2524 and two\nknown AGNs to suggest that while the 1--5 \\hbox{\\micron~} emission in IRAS\n05189$-$2524 may be due to hot dust heated by an AGN, the 8--100\n\\hbox{\\micron~} emission may be better explained as arising from star\nformation traced by the observed PAH emission. In Fig. 2\nthe power-law extrapolations of the dust continuum emission\n(solid lines) found in the 8--13 \\hbox{\\micron~} spectra are projected to 5\nand 25 \\hbox{$\\mu$m}. For the two sources where the 8--13 micron\nemission is thought to be due to AGN-heated dust (Mkn 231 and\nIRAS 13349+2438) (Roche et al. 1983; Beichman et al. 1986),\nthese extrapolations are reasonable predictors of the observed\nflux densities at both 5 and 25 \\hbox{$\\mu$m}. For IRAS 05189$-$2524\nthe extrapolation of the continuum to 25 \\hbox{\\micron~} is in reasonable\nagreement with the observed flux density; however the\nextrapolation to 5 \\hbox{\\micron~} fails to predict the 4--5 \\hbox{\\micron~}\nphotometry of IRAS 05189$-$2524 (see double arrow), suggesting\nthat an unrelated power source is responsible for this\nemission. Given the apparent similarity in the shape of 1--5\n\\hbox{\\micron~} emission in both IRAS 05189$-$2524 and IRAS 13349+2438, it\nis possible that the spectral energy distribution of IRAS\n05189$-$2524 is the result of a starburst producing the majority\nof the 8--100 \\hbox{\\micron~} emission and an AGN with a spectral energy\ndistribution similar to that of IRAS 13349+2438 responsible\nfor the bulk of the shorter wavelength emission but\ncontributing little to the 8--100 \\hbox{\\micron~} emission. Such a\ndecomposition could reconcile the apparently contradictory\nobservations of both an AGN-like optical spectrum and the\npresent PAH type 8--13 \\hbox{\\micron~} spectrum. A decomposition of this\nkind could also be consistent with the suggestion of Goldader\net al. (1995) that the 2 \\hbox{\\micron~} continuum is dominated by AGN\n[heated dust] emission based on the small equivalent width of the observed\n\\hbox{Br$\\gamma$}~ emission. That hot dust typically plays a lesser\nrole in starburst galaxies at 2 \\hbox{\\micron~} may also be deduced\nfrom the frequent detection of stellar continuum with probably\nlittle veiling of the CO absorption features, while in IRAS\n05189$-$2524 CO absorption is not detected (Goldader et\nal. 1995). Another piece of evidence that supports the view that\nthe AGN in IRAS 05189$-$2524 may not be energetically dominant\nis the failure to detect a radio core of extreme brightness\ntemperature in VLBI observations (Lonsdale, et al.\n1993).\n\nThe VLA radio map of CHYT shows the radio continuum in this\nsource to be extended on a scale of 160 pc, which is\ncomparable to the minimum blackbody diameter of 280 pc based on\nthe observed 60 and 100 \\hbox{\\micron~} flux densities and assuming optically thick\ndust emission. It is therefore plausible that the far-infrared\nemission arises from optically thick warm dense clouds embedded\nin the starburst with a low volume filling factor but a column\ncovering factor of order unity. The mid-infrared continuum and\nPAH emission present in this spectrum might arise from the\nsurfaces of these clouds and may be optically thin in the sense\nthat the emission arises from cloud surfaces facing our line of\nsight, and from a layer that is only 1--2 $A_V$ thick on any\ngiven line of sight into the starburst. \n\nGiven the possibility\nof a decomposition of the 1--100 \\hbox{\\micron~} spectral energy\ndistribution into a dominant starburst component and an AGN\ncomponent important only at shorter wavelengths, the\nplausibility of the infrared emission being spatially associated\nwith the radio continuum, and most importantly, the presence of\n11.3 \\hbox{\\micron~} PAH emission in the 8--13 \\hbox{\\micron~} spectrum taken together with the\nadequate fit of the model given in Fig. 1, the 8--1000 \\hbox{\\micron~}\nemission from IRAS 05189$-$2524 is attributed to a starburst for\nthe purposes of this work.\n\n\\subsubsection{Digression on newly available {\\em ISO} data}\n\nThe model applied to the data presented in Fig. 1 for the\npurpose of spectral classification is not required in cases\nwhere the strength of the 7.7 \\hbox{\\micron~} PAH feature has been\nobserved as is now the case for four of the ultraluminous\ninfrared galaxies in this combined sample as a result of {\\em\nInfrared Space Observatory (ISO)} observations (Genzel et\nal. 1998). With the additional information it is possible to\ncompare observations with a variety of scenarios to better\nunderstand the physical situation with greater confidence than\npossible when only ground-based data are available. In Fig. 3,\nthree possibilities are considered: (1) For the top spectrum\nin each panel, a PAH plus power-law continuum model such as\nthat employed in Fig. 1 is fit not just in the spectral region\nsurrounding the 11.3 \\hbox{\\micron~} feature, but to the entire 6--13 \\hbox{\\micron~}\nrange. (2) For the middle spectrum in each panel, a PAH plus\npower-law continuum model is subjected to extinction by cold\ndust that includes the silicate absorption feature, as adopted\nby Dudley \\& Wynn-Williams (1997), and again fit to the entire\n6--13 \\hbox{\\micron~} range. (3) For the bottom spectrum in each panel, a\nPAH plus power-law continuum model is additively combined with\na silicate-absorbed continuum source, and again fit over the entire\n6--13 \\hbox{\\micron~} range. The model fits presented in Fig. 3 have been\ncarried out using CURVEFIT with equal weighting. In the first\nmodel, the index of the power-law was specified either from\nTable 3 (IRAS 17208$-$0014 and Mkn 273) or as 12.8 and 4.9 for\nArp 220 and Mkn 231, respectively. It should be noted that\nthree of the sources exhibit a large dynamic range, and in the\nfollowing discussion some attention will be drawn to the poor\nfit at low flux density levels, which is also an expected systematic\neffect of the fitting procedure, since differences between\nsmall numbers are small. It may also be noted that the\nmismatches are apparent for a large number of neighboring data\npoints.\n\n\n\\begin{figure}\n\\end{figure}\n\n\n\n\n\\subsubsection{IRAS 17208$-$0014}\n\nIn IRAS 17208$-$0014, all models give fairly adequate fits\nto the data, and deciding between them depends mainly on\nother observational evidence. The strongest evidence that\nsupports a preference for the fit shown for the top spectrum\nis that this source is extended on a scale $\\sim$1.7 kpc in\nthe radio (Condon et al. 1996), although higher spatial\nresolution observations reveal the presence of a sub-kpc\nradio continuum source associated with the OH maser emission\n(Martin et al. 1989). The failure of Martin et al. (1989)\nto detect more extended emission may be due to the\nrestricted set of base lines they employed. When this\nevidence is taken along with the optical spectral\nclassification of Veilleux et al. (1995) in which the line\nratios in this source are similar to those in \\ion{H}{2}\nregions, there is no strong reason to suspect that anything\nother than star formation is responsible for the observed\nmid-infrared spectrum. Nor is there any reason to suspect a\nhigh degree of extinction; IRAS 17208$-$0014 is among the\nsources presented by Genzel et al. (1998) for which the\ncontinuum underlying the 12.8-\\hbox{\\micron~} [\\ion{Ne}{2}] feature in\ntheir SWS data can be comfotably plotted on the scale of the\nISOPHOT-S data in their fig. 2, suggesting a lack of strong\nsilicate absorption (see also Fig. 1). The fit to the upper\nspectrum might be improved somewhat by increasing the\nstrength of the 7.7 and 6.2 \\hbox{\\micron~} features by a factor of 2,\na procedure which is probably allowed for the 7.7 \\hbox{\\micron~}\nfeature at least since the Orion Bar spectrum used as a\ntemplate here is known to have a somewhat weaker than usual\n7.7 \\hbox{\\micron~} feature relative to the 11.3 \\hbox{\\micron~} feature (Zavagno\net al. 1992).\n\n\\subsubsection{Mkn 273}\n\nThe fit to the top spectrum of Fig. 3 for Mkn 273 gives, as\nmight have been anticipated from Fig. 1, a poor representation\nof the data. The fitted strength of the 11.3 \\hbox{\\micron~} feature is\nclearly too large, which is roughly the converse of the result\nin Fig. 1. Thus, it may be concluded that a simple PAH plus\npower-law continuum model does not describe these data. The fit\nshown for the middle spectrum is better, in that it requires\nless PAH emission over all and the strength of the 11.3 \\hbox{\\micron~}\nfeature is not over-predicted; however the continuum around the\n9.7 \\hbox{\\micron~} silicate minimum is not well reproduced. This\nsuggests that an unaccounted-for emission component is required\nto produce this emission. The fit to the lower spectrum\novercomes this difficulty by superimposing an unobscured PAH plus power-law\ncontinuum such as that which fit IRAS 17208$-$0014 (top\nspectrum) upon a continuum source that suffers strong silicate\nabsorption. Since Mkn 273 shows clear signs of nuclear\nactivity (Armus, Heckman \\& Miley 1989; Veilleux et al. 1995),\na silicate-absorbed continuum source might be produced by an\nedge-on view of a disk model such as those proposed by Pier \\&\nKrolik (1992) or Efstathiou \\& Rowan-Robinson (1995), which are\nnatural elaborations of the AGN unification model proposed by\nAntonucci \\& Miller (1985). The continuum level evident in the\nSWS observations of 12.8-\\hbox{\\micron~} [\\ion{Ne}{2}] in Genzel et\nal. (1998) is consistent with the new CGS3 data presented here;\nand therefore consistent with silicate absorption playing a\nrole in this source. If the fit to the bottom spectrum is accepted,\nthen the AGN component could well be responsible for the\ngreater fraction of the 60 and 100 \\hbox{\\micron~} emission.\n\n\\subsubsection{Arp 220}\n\nPrior to the new {\\em ISO} observations the analysis of the\nArp 220 spectrum would have proceeded very much along the\nlines that have just been given for Mkn 273 (Smith et\nal. 1989), with the proviso that Arp 220 has less obvious\nsigns of activity in its nucleus so that disk-like models\nfor the silicate-absorbed continuum emission may be less\nhelpful in advancing our understanding, and deeply embedded\nmodels perhaps more descriptive (Dudley \\& Wynn-Williams\n1997). Applying the model used in Fig. 1 to the data of\nSmith et al. (1989) severely under-predicts the observed 8.2\n\\hbox{\\micron~} flux density. However, as can be seen in Fig. 3 the\nnew {\\em ISO} data can apparently be fit by a simple PAH\nplus power-law continuum model due to the apparent\ndiscrepancy between the observations of the 11.3 \\hbox{\\micron~}\nfeature reported by Smith et al. (1989) and the {\\em ISO}\ndata of Genzel et al. (1998). It should be noted that there\nare no continuum points longward of the 11.3 \\hbox{\\micron~} feature in\nthe new {\\em ISO} data, so that determining its strength\nfrom those data is difficult. It seems safer at this point\nto accept the ground-based data as a guide in this spectral\nregion and thus hold this model in abeyance, pending further\nobservational evidence. The fit to the middle spectrum\nis essentially the same as that given in fig. 1 of Smith et\nal. (1989) and suffers from the same difficulty noted by\nSmith et al. (1989), namely that the minimum near 9.7 \\hbox{\\micron~}\nis poorly fit. The fit to the lower plot is similar to that\nshown in Smith et al.'s (1989) fig. 2 or as adopted by Dudley \\&\nWynn-Williams (1997). The continuum in the 12.8-\\hbox{\\micron~}\n[\\ion{Ne}{2}] SWS data of Genzel et al. (1998) is consistent\nwith the Smith et al. (1989) data but is not apparently\nconsistent with the ISOCAM-CVF spectrum presented by Elbaz\net al. (1998).\n\nFor Arp 220, therefore, the new {\\em ISO} data presented by\nGenzel et al. (1998) seems to allow the possibility that\nunextinguished PAH emission plus a simple power-law\ncontinuum describes this spectrum as a result of the\napparent strength of the 11.3 \\hbox{\\micron~} feature in the new data.\nIt has been suggested that differing spectral resolution\n(roughly a factor of 2) between the Smith et al. (1989) and\nGenzel et al. (1998) data may account for this difference,\nhowever this could only work and if the 11.3 \\hbox{\\micron~} feature\nwere unresolved in both spectra, and that does not appear to\nbe the case. Gauging the over all PAH emission from the 6.2\n\\hbox{\\micron~} feature which displays continuum on either side, the\nintensity of which is found to be well correlated with the\nthat of the 7.7 \\hbox{\\micron~} feature in Galactic sources (Cohen\net al. 1989), it is clear the PAH emission relative to the\nover-all 8--1000 \\hbox{\\micron~} emission is exceedingly weak, really\nno stronger than what might be allowed in Mkn 231 (see\nnext), if the apparent 6.2 \\hbox{\\micron~} feature in the {\\em ISO}\ndata for Mkn 231 is real. In contrast, again using the 6.2\n\\hbox{\\micron~} PAH feature as a guide, the strength of the PAH\nemission relative to the 8--1000 \\hbox{\\micron~} emission in Mkn 273 is\ntwice as strong, and in IRAS 17208$-$0014, five times as\nstrong. The Arp 220 spectrum might possibly be produced by a\nstarburst that reprocesses more that 90 per cent of its\nluminosity not once but many times so that the observed PAH\nemission is produced only in a thin outer shell or in a\nsmall number of narrow chinks. Such a model would be even\nmore extreme than what is argued above as plausible for IRAS\n05189$-$2524 in a starburst context. However, given the\ndisagreement between the two spectra of this source near\n11.3 \\hbox{\\micron~} and in consideration of the arguments of Dudley \\&\nWynn-Williams (1997), a model that has both unobscured PAH\nemission and a silicate-absorbed continuum due to a deeply\nembedded AGN continues to seem more natural. This\ninterpretation also seems to account simultaneously for both\nthe 158-\\hbox{\\micron~} [CII] (Luhman et al. 1998) and 2-10 keV X-ray\n(Iwasawa 1999) deficits in this source if the ``AGNs''\n(Soifer et al. 1999) are sufficiently obscured. It should\nbe noted that in this preferred model, only a fraction of\nthe flux density at 7.7 \\hbox{\\micron~} can be attributed to the 7.7\n\\hbox{\\micron~} PAH feature, in accordance with either the 12.8 of 14.3\n\\hbox{\\micron~} continuum seen in the SWS data presented by Genzel et\nal. (1998). As suggested by Dudley \\& Wynn-Williams (1997),\nspatially resolved spectroscopic observations may help to\nclarify this issue.\n\n\\subsubsection{Mkn 231}\n\nFor Mkn 231 any PAH contribution to the spectrum must be\nminimal compared to a continuum contribution (Roche et\nal. 1983), as can also be seen from the fit to the upper\nspectrum in Fig. 3. This source had been fit adequately\nwith a model consisting of a continuum source suffering\nsilicate absorption by Roche et al. (1983), and the present\nmodels are {\\it not} an improvement. It would be of\ninterest to know if disk radiative transfer models such as\nthose proposed by Pier and Krolik (1992) would fair as well\nas the cold absorber model of Roche et al. (1983) in the\ndetailed fit to the silicate absorption feature given their\nlikely success in fitting the overall infrared spectral\nenergy distribution of Mkn 231. The 11.3 \\hbox{\\micron~} emission\nfeature is not detected in this source and it is not clear\nthat the single point that is spectrally coincident with the\n6.2 \\hbox{\\micron~} feature in the {\\em ISO} data should be considered\nas a real indication of the presence of PAH emission. Some\nPAH emission is not unexpected in this source given the\npresence of star-forming knots near the nucleus of this\nsource that might account for $\\sim$10 per cent of its\ninfrared luminosity (Surace et al. 1998). The estimate of\nthe 7.7 \\hbox{\\micron~} PAH strength given by Genzel et al. (1998)\nrelies on a continuum point at 10.9 \\hbox{\\micron~} which lies in the\nsilicate absorption feature. Of the sources measured in\nthis way by Genzel et al. (1998) Mkn 231 perhaps provides\nthe clearest example of how this method is prone to\noverestimate the PAH strength when silicate absorption is\npresent. The 12.8 or 14.3 \\hbox{\\micron~} continuum in their fig. 1\nprovide a rough guide as to when there may be a problem in\nthis respect. For Mkn 231, the 12.8 \\hbox{\\micron~} continuum of\nGenzel et al. (1998) is high in comparison with the data of\nRoche et al. (1983) but it is possible that it would agree\nwith the extrapolation of the long-wavelength ISOPHOT-S data\nwhich also apparently fall significantly above the Roche et\nal. (1983) data as seen in Fig. 3. It is not clear if these\ndifferences should be attributed to aperture effects, or\nother systematics, or if they reflect intrinsic source\nvariability over 1.4 decade; Roche et al. (1983) found good\nagreement with the narrowband photometric observations of\nReike (1976) made over 0.4 decade earlier.\n\n\n\\subsection{A 60 \\protect\\hbox{$\\mu$m}~ flux-limited subsample\\label{ch5:d2}}\n\n\n\nAmong galaxies with \\hbox{$L_{\\rm IR}$} $\\geq 1.6\\times10^{11}$ \\hbox{L$_\\odot$}, a nearly\ncomplete (all but three sources have 8--13 \\hbox{\\micron~} spectra)\nflux-limited (at 60 \\hbox{$\\mu$m}) sample of 25 sources may now be\ndefined on the combined Bright Galaxy Survey (Soifer et\nal. 1989) and the Extended Bright Galaxy Survey (Sanders et\nal. 1995) (combined BGS) for $F_\\nu$(60) $\\geq$ 11.25 Jy\n(roughly twice the flux limit of combined BGS). The only other\nrestriction on this subsample is that the sources be observable\nusing UKIRT, i.e. $-25^\\circ$ $\\leq \\delta \\leq $ 60$^\\circ$.\nTable 5 gives the ultraluminous and luminous samples with the\nsource name in column 1, the 8--13 spectral classification in\ncolumn 2, the reference for the 8--13 \\hbox{\\micron~} spectrum in column 3,\nthe $\\log$ of the 8--1000 \\hbox{\\micron~} luminosity in column 4, and the\n$\\log$ of the ratio of the 12 \\hbox{\\micron~} flux density to the 60 \\hbox{\\micron~}\nflux density in column 5. The ultraluminous sample has 5 members\nand the luminous comparison sample has 20 members. The\nultraluminous sample is {\\em unlikely} to span all the possible\ncharacteristics of ultraluminous infrared galaxies and therefore\nis not fully representative. There are two observations that\nconfirm this: (1) if Arp 220 were scaled to the 60 \\hbox{\\micron~} 11.25 Jy\nflux limit used to define this sample, it would not have been\nobservable using CGS3 or other ground-based spectrographs\navailable when these data were collected, and (2) the ratio\ngiven in column 5 of Table 5 for the two brightest ultraluminous\ngalaxies (Arp 220 and Mkn 231) bracket the same ratio for all\nthe galaxies in the sample with $1.6\\times10^{11} \\leq$ \\hbox{$L_{\\rm IR}$}\n$\\leq 10^{12}$ \\hbox{L$_\\odot$}~ except NGC 1068 and NGC 7469. An analysis\nover the entire Bright Galaxy Survey shows that the dispersion\nof this ratio is indeed larger for ultraluminous infrared\ngalaxies than for luminous infrared galaxies by a factor of 2\n(Soifer \\& Neugebauer 1991). Thus, conclusions based on the\npresent flux-limited sample of ultraluminous galaxies should be\nviewed with caution not simply because of the small number of\ngalaxies in this sample (counting statistics on a bimodal parent population), but because the\ncharacteristics of the parent population are not well\nrepresented (under-sampling).\n\n\n\nIt is possible that the last of these two difficulties\ncould be somewhat ameliorated by the following\nconsiderations. Ultraluminous galaxies that fall outside\nthe range of 12 on 60 \\hbox{\\micron~} flux density ratios defined by\nArp 220 and Mkn 231 may be assigned to the AGN-dominated\nbin in the context of exploring the AGN vs. starburst\nhypotheses. For ultraluminous galaxies with 12\/60 \\hbox{\\micron~}\ncolors redder than Arp 220, PAH emission cannot make a\ncontribution to the 12 \\hbox{\\micron~} flux density that is\nsufficiently large to explain the 60 \\hbox{\\micron~} flux density. \nIt is therefore likely that such sources will have deep\nsilicate absorption features, and by the arguments\npresented by Dudley \\& Wynn-Williams (1997), the size of\ntheir power source should be too small to be due to a\nstarburst. For galaxies with 12\/60 \\hbox{\\micron~} colors bluer than\nMkn 231, 60 \\hbox{\\micron~} selected samples will overlap 25 or 12\n\\hbox{\\micron~} selected samples that are known to select for AGNs\n(de Grijp et al. 1985; Spinoglio \\& Malkan 1989). However,\nreliance on these notions would be ill advised given the\ninternal evidence in the present data showing that PAH\nemission can be important in ultraluminous galaxies with\n12 to 60 \\hbox{\\micron~} flux density ratios similar to the two\nextremes, namely, IRAS 17208$-$0014 and IRAS 05189$-$2524.\n\n\n\n\n\n\nInspection of Table 5 shows that 3 out of 5 of the ultraluminous\ngalaxies are powered predominantly by AGNs although if judgment\nis reserved for Arp 220, this fraction would be 2 out of 5. The\ntrend is thus consistent with about half of ultraluminous\ngalaxies being powered by AGNs. On the other hand, the fraction\nof the more numerous luminous galaxies in this sample that are\npowered predominantly by AGNs is at most 0.25 and could be as\nsmall as 0.05 or smaller if the relative contributions of the\nAGN and starburst in NGC 1068 are approximately equal (Telesco\net al. 1984). This trend supports the evidence from optical\nspectroscopy that AGNs are present in greater numbers in\nultraluminous infrared galaxies (Veilleux et al. 1995). It goes\nfurther than optical or radio studies because it is possible to\ndemonstrate with greater certainty which type of power source\n(starburst or AGN) dominates the infrared luminosity. Further,\nit is apparent that for luminosities less than 10$^{12}$ \\hbox{L$_\\odot$},\nstar formation is far and away the dominant power source, a\nresult that confirms work at many wavelengths. This analysis\nlends some support to the proposal of Sanders et al. (1988) that\nthere is an evolutionary connection between ultraluminous\ninfrared galaxies and quasars, while at the same time\ndemonstrating that massive star formation can indeed be the\ndominant power source of some ultraluminous infrared galaxies.\nHowever, it must be conceded that the evidence presented\nconstitutes only a trend. The precise proportion of\nAGN-dominated sources among 60 \\hbox{\\micron~} selected ultraluminous\ninfrared galaxies cannot be said to have been established here,\nand thus no statistically meaningful difference between luminous\nand ultraluminous galaxies has truly been shown. It is\nof interest that the trend does continue with the next step down\nin 60 \\hbox{\\micron~} flux density. Adding IRAS 12112+0305 and IRAS\n08572+3915 to the present sample of five ultraluminous galaxies adds\none apparently starburst-dominated (based on the apparently\nnormal ratio of the 6.2 \\hbox{\\micron~} PAH feature emission to overall infrared\nemission in the IRAS 12112+0305 spectrum of Genzel et\nal. 1998\\footnote{A crude attempt to correct for the\ncontribution of the silicate absorption edge to the 7.7 \\hbox{\\micron~} PAH\nfeature by fitting a power law to continuum points at 5.9 and\n14.3 \\hbox{\\micron~} suggests a deficit in the strength of the 7.7 \\hbox{\\micron~} feature for\nIRAS 12112+0305.})\nand one AGN-dominated object (Dudley \\& Wynn-Williams 1997).\n\n\\subsection{Anti-correlation between 11.3 \\hbox{\\micron~} eqivalent width and\n$L_{\\rm IR}$\/$M({\\rm H}_2)$\\label{ch5:d3}}\n\n\nIn Fig. 4 a possible anti-correlation between\nthe equivalent width (EW) of the 11.3 \\hbox{\\micron~} feature and the global ratio of $L_{\\rm IR}$\/$M({\\rm H}_2)$ is presented for a number of galaxies. The\nEW data are based either on the\nmodel fits given in Table 3 or the previously published\nvalues of RASW reproduced in Table 6. The ratios of\ninfrared luminosity to H$_2$ mass are taken from the\nliterature (Tables 4 and 6). The solid\nline in Fig. 4 indicates inverse proportionality.\n\n\n\n\n\nThe ratio ${L_{\\rm IR} \\over {\\rm M(H}_2)}$ is often taken to be\na measure of star formation efficiency. When it is large, then\na large fraction of the available gas is involved in star\nformation; when it is small, conditions may be such that the gas\nexists in a more quiescent state. However, for a very large\nratio of ${L_{\\rm IR} \\over {\\rm M(H}_2)}$, Sanders et\nal. (1991) argued that even a very high star formation\nefficiency cannot account for the infrared luminosity, on the\ngrounds that star formation would too quickly consume the\navailable gas mass, so that AGN-like activity might be the more\nlikely primary power source of the infrared luminosity. Both\n${L_{\\rm IR} \\over {\\rm M(H}_2)}$ and EW(11.3) are independent\nof distance in so far as EW(11.3) is independent of aperture\nsize.\n\nA tendency for EW(11.3) to be anti-correlated with the star\nformation efficiency might arise if two conditions were met.\nFirst, if the intensity of the 11.3 \\hbox{\\micron~} feature is a roughly\nconstant fraction of the overall infrared emission (proportional\nto $L_{\\rm IR}$) then this will not contribute to the\nanti-correlation. Second, if the strength of the continuum at\n11.3 \\hbox{\\micron~} increases with decreasing availability of dust\n(inversely proportional to M(H$_2$)) to reprocess the light\nemitted by massive stars so that a larger fraction of the dust\ngrain population is heated to the warm temperatures required for\nthe dust grains to emit at 11.3 \\hbox{\\micron~} then this could be the\nsource of anti-correlation. Such an explanation requires\nfurther elaboration since the process by which the 11.3 \\hbox{\\micron~}\ncontinuum is produced is likely, in general, to be a local\nphenomenon rather than a global one.\n\nAn elaboration that may be in keeping with recent {\\em HST}\nobservations (Watson et al. 1996; Surace et al. 1998;\nSatyapal et al. 1997) is that as the star formation\nefficiency increases, the proportion of massive stars formed\nin dense clusters may increase. If this is the case, then\nthe time-scale for clearing the interstellar medium from\naround these clusters to a distance where dust grains would\nno longer be heated sufficiently to contribute to the 11.3\n\\hbox{\\micron~} continuum may be longer that the time-scale for\nclearing the interstellar medium around the groups of\nmassive stars that might form if the star formation\nefficiency is lower. Two effects may contribute to this.\nFirst, the distance at which a grain could be heated to a\ntemperature high enough to contribute to the 11.3 \\hbox{\\micron~}\ncontinuum would increase as the square root of the number of\nmassive stars in a cluster so that if the velocity of gas\nand dust being cleared from clusters of massive stars is not\ntoo different from the velocity of gas and dust cleared from\ngroups of massive stars, the dust remains heated for a\nlonger period of time. Second, dense clusters might be\ntemporarily gravitationally bound and behave ballistically\nwith respect to the dissipative gas. They may sometimes move\ninto a fresh reservoir of gas, heating the associated dust\nto higher temperatures.\n\nIn this view, the additional contribution to the 11.3 \\hbox{\\micron~}\ncontinuum emission at higher star formation efficiency would\nbe provided by possibly equilibrium temperature grains\nheated by a combination of direct and Ly$\\alpha$ heating\nrather than small grains heated by a single photon. To\nzeroth order, this is a sensible proposition, since one\nexpects (at least within photo-dissociation regions) that\nthe PAH emission and the spike-heated grain emission to have\nroughly fixed relative contributions at 11.3 \\hbox{\\micron~} due to the\nrough similarity of the emission time scales for the two\nemission processes, so that the apparent anti-correlation\nwould not necessarily arise.\n\nA number of observations will be required to examine this\nsituation. Most important are higher signal-to-noise\nobservations that will allow real measurements of EW(11.3)\nfor a number of the sources included in Fig. 4. These are\nrequired to decide if the anti-correlation is has any merit.\nAnother avenue of investigation would be to make higher\nspatial resolution observations of nearby starbursts where\nclusters of massive stars are observed to exist to establish\nwhether or not EW(11.3) does decrease in the vicinity of\nsome clusters (see Dudley \\& Wynn-Williams 1999 for a\npreliminary attempt). Finally, high spatial resolution 2\n\\hbox{\\micron~} adaptive optics imaging of a number of starbursts with\na range of star formation efficiency can help to determine\nif the fraction of massive stars that exist in dense\nclusters varies with the star formation efficency parameter\nas conjectured here.\n\n\\section{Conclusions\\label{ch5:c}}\n\nNew 8--13 \\hbox{\\micron~} spectroscopy of 27 infrared galaxies has been\npresented. The spectra have been fit with a two-component model\nconsisting of a power-law continuum underlying a PAH emission\nspectrum. These fits are based on the spectral data between 10 and\n13 \\hbox{\\micron~} and are used to predict the flux\nemitted between 8.0 and 8.4 \\hbox{$\\mu$m}. It is\nfound that for the majority of the observed galaxies, the\nmodel predicts the observed 8.0--8.4 \\hbox{\\micron~} emission to\nwithin an acceptable factor of 2, suggesting that the\nmajority of the sources are powered by starbursts.\n\nOne source, Mkn 273, has been found to show very clear evidence for\nsilicate absorption.\n\nA flux-limited 60 \\hbox{\\micron~} selected subsample that combines the\npresent spectroscopic results and those of RASW had been\ndefined, consisting of 25 galaxies. In this subsample the fraction of ultraluminous\ninfrared galaxies powered by AGN is $\\sim$50 per cent in\ncontrast to 5--25\\% among the luminous infrared galaxies. \nOwing to the small sample size for the ultraluminous infrared\ngalaxies, this difference in AGN fraction can be considered only\na trend.\n\nA possible anti-correlation is identified between the\nequivalent width of the 11.3 \\hbox{\\micron~} PAH feature and the star\nformation efficiency in galaxies where both have been\nmeasured. It is suggested that variation in the continuum\ndust temperature as a function of star formation efficiency\ncould produce this apparent effect.\n\n\n\n\\noindent{ACKNOWLEDGMENTS}\n\nThe author would like to thank the UKIRT staff: Joel Aikock,\nTim Caroll, Dolores Walthers, and Thor Wold, for assistance at\nthe telescope, and John Davies, Tom Geballe, and Gillian Wright\nfor fruitful discussions and calibration data in advance of\npublication. UKIRT is operated by the Joint Astronomy Centre\non behalf of the U.~K.~Particle Physics and Astronomy Research\nCouncil. Brian Rush, Reinhart Genzel, Vassilis Chamanadaris,\nJoe Hora, and D.-C. Kim all kindly provided data in advance of\npublication. A series of exchanges with Dieter Lutz has help to\nclarify certain issues, as has a discussion with Gary\nNeugebauer, Keith Matthews and Eichi Egami. Eric Becklin has\nprovided thoughtful comments at several stages of this work.\nMany members of the Institute for Astronomy have benefited this\nwork through discussion and sharing of work in progress. A few\nare Gareth Wynn-Williams (dissertation committee chair), Dave Sanders, Bob Joseph, Klaus\nHoddap, Josh Barnes, Alan Tokunaga, Jeff Goldader, Jason\nSurace, and Masatoshi Imanishi; thanks are due to these and\nothers. Louise Good proofread the manuscript prior to\nsubmission. This work has been partially supported by NSF\ngrants ASTR-8919563, NASA grant NAGW-3938, and the Frank Orrall\nfund. The NASA Extragalactic Database (NED) has been a\nconstant aid. It is run by the Jet Propulsion Laboratory,\nCalifornia Institute of Technology, under contract with NASA.\nThis work has also made use of NASA's Astrophysics Data System\nAbstract Service.\n\n\\newpage\n\n\n","meta":{"redpajama_set_name":"RedPajamaArXiv"}} +{"text":"\\section{Introduction}\n\\label{intro}\n\nUnderstanding the interplay of rare-earth local-moment magnetism\nand $3d$ transition metal itinerant magnetism is of fundamental\nphysical interest\\cite{Campbell72} and may help in the design of\nmore efficient permanent magnets.\nThe tetragonal $R$Fe$_4$Al$_8$ ($R=$rare earth) compounds\n(Fig.\\ \\ref{Fig1}) are well suited for studying this interplay\nbecause of simple symmetry conditions and because the interaction\nbetween the two magnetic sublattices is rather weak: Fe moments\nappear to order without a corresponding $R$ $4f$ order.\nConsequently, numerous studies have been performed on\n$R$Fe$_4$Al$_8$ (R148) over the last 30 years and interest in\nthese compounds has remained\nhigh.\\cite{Buschow78,Felner78,Fujiwara87,Duong01,Duong02,Langridge99,Paixao00,Paixao01,Schobinger98,Schobinger99}\n\n\\begin{figure}[tb]\n\\includegraphics[width=0.98\\linewidth]{pixGd148struc.eps}\n\\caption{(Color online) The crystal structure of $R$Fe$_4$Al$_8$.\nThin lines outline the tetragonal (I$4\/$mmm) unit cell, thick\nlines denote the Fe ``cage'' around the rare earth $R$. }\n\\label{Fig1}\n\\end{figure}\n\n\nNeutron scattering studies performed on R148 with various $R$\nindicate that Fe moments, between which the coupling is strongest,\norder between $100$ and $200\\, {\\rm K}$, generally in cycloidal\nstructures with propagation vectors parallel to [110] and moments\nconfined to the $a$$-$$b$\nplane.\\cite{Paixao00,Paixao01,Schobinger98,Schobinger99} For\nmagnetic rare earths $R$, ordering of the $R$ moments, with much\nweaker coupling and at much lower temperature $T$, has been\nreported. However, the critical temperature associated with this\nordering is very poorly defined. For example in Dy148 at\n$\\sim$$50\\,{\\rm K}$ the magnetic dc susceptibility rises faster\nthan expected for Curie-Weiss behavior, but there are no sharp\nfeatures.\\cite{Paixao00} Furthermore, the rare earth moments were\nfound to follow the Fe-moment\nmodulation.\\cite{Paixao00,Schobinger98} This seems to imply that\nthe Fe$-R$ moment interaction is stronger than the $R$$-$$R$\nmoment interaction.\n\nOf all rare earth elements, Gd has two highly interesting specific\nfeatures associated with its high-spin and zero-angular-momentum\n$4f$ state: 1) high spin results in strong magnetic coupling\nbetween the localized $4f$ and the conduction electrons, implying\nlarge magnetic interactions and the largest de Gennes factor of\nall rare earths; 2) zero angular momentum implies a spherical $4f$\ncharge-cloud and no magneto-crystalline anisotropy (MCA) resulting\nfrom crystal-electric-field effects $-$ the direct interplay of\nmagnetic interactions can be studied without\ncrystal-electric-field effects. In view of this, it is somewhat\nsurprising that GdFe$_4$Al$_8$ (Gd148) has been much less\nstudied\\cite{Fujiwara87,Duong01} than R148 with other rare earths.\nIn particular, apart from M{\\\"{o}}ssbauer\nspectroscopy,\\cite{Buschow78} there have been no microscopic\nstudies of Gd148.\n\nHere, we report on the flux-growth of single crystals of the\nGdFe$_4$Al$_8$ phase, and on the characterization of the crystals\nby magnetization, electrical transport, specific heat, and x-ray\ndiffraction. We also report on first synchrotron x-ray resonant\nmagnetic scattering (XRMS) data. We provide evidence for two\nconsecutive phase transitions at low temperature $T$ in addition\nto the N\\'{e}el transition at $155\\, {\\rm K}$, and show that the\ntwo, low-$T$, transitions are associated with the ordering of the\nGd $4f$ moments, resulting in complex magnetic structures\ninvolving both ferro- and antiferromagnetic components. In\ncontrast to other R148 compounds, the $4f$-moment ordering is\nassociated with a propagation vector that is distinct from that\nassociated with the Fe moment order, which at low $T$ has the same\nmodulation as in the vicinity of the {N{\\'{e}}el} temperature. We\nalso present the magnetic phase diagram for two field directions\nand discuss the implications of changes in the Fe stoichiometry.\n\nThe paper is organized into seven sections. In Sec.\\ \\ref{exp} we\ndescribe the flux growth of the single crystalline samples used in\nthe study, and the experimental procedures. In Sec.\\ \\ref{lowH}\nzero and low field electrical transport, magnetization, and\nspecific-heat data are presented, and we evaluate the influence of\nthe iron stoichiometry on the physical properties. In Sec.\\\n\\ref{highH} we present electrical transport and magnetization data\nmeasured in fields for two field directions, and in Sec.\\\n\\ref{xrms} the results of a first XRMS experiment are described.\nFinally, we discuss the low-temperature magnetic phase diagram and\npossible magnetic structures in Sec.\\ \\ref{disc}, before\nsummarizing our main conclusions in Sec.\\ \\ref{conc}.\n\n\\section{Experimental}\n\\label{exp}\n\nWhereas all previous single crystals of $R$Fe$_4$Al$_8$ material\nwere obtained using the Czochralsky-method, we have grown single\ncrystals of Gd148 (and other R148) with a self-flux\nmethod.\\cite{Sol1,Sol2,Sol3} One of the problems with studies on\nsingle crystals of R148 is a width of formation often observed,\ninvolving some Fe atoms occupying nominal Al positions and vice\nversa; this can lead to striking differences in the magnetic\nstructure and phase diagram as compared to the stoichiometric\nmaterial (see, e.g., Refs.\\ \\onlinecite{Paixao01,Waerenborgh00}).\nA flux-growth procedure may allow better control of the\nstoichiometry of the crystals by varying the starting composition.\nUnfortunately, information about the ternary phase diagram\n$R$-Fe-Al is very limited. We used differential thermal analysis\n(DTA; see Ref.\\ \\onlinecite{Janssen05} for a review of the use of\nDTA in the flux growth of crystals) to establish i) that\nGdFe$_4$Al$_8$ is congruently melting and ii) selected solidus\ntemperatures in the ternary around the GdFe$_4$Al$_8$\nstoichiometry. For the crystals used in this study, we selected\nstarting compositions between Gd148 and the also congruently\nmelting Fe$_2$Al$_5$ (Gd$_3$Fe$_{16}$Al$_{34}$ and\nGd$_2$Fe$_{14}$Al$_{31}$), which should not lead to large\nimbalances in the Fe to Al ratio.\n\nElements in ratios corresponding to the starting composition were\nfirst arc-melted together. The resulting ingot was placed in an\nalumina crucible, which was wrapped in Ta foil (to prevent any\nresidual oxygen in the Ar from reacting with the sample) and\nplaced in a vertical tube furnace in flowing Ar. Crystals were\ngrown by heating to $1475^{\\circ}{\\rm C}$ and then slowly\n($2^{\\circ}{\\rm C} \/ {\\rm h}$) cooling to $1180^{\\circ}{\\rm C}$,\nat which point the furnace was turned off. The flux was removed\nfrom the crystals in a second step by heating to $1200^{\\circ}{\\rm\nC}$, keeping the temperature for $30\\,{\\rm min}$ and then\ndecanting, following procedures described in Refs.\\\n\\onlinecite{Sol1,Sol2,Sol3,Janssen05}.\n\nFor both starting compositions, we obtained crystals of the Gd148\nphase, as identified via powder x-ray diffraction. Crystals\ntypically grow prismatically with (110) facets and the long\ndirection parallel to [001], as determined with Laue scattering.\nWe obtained crystals of dimensions up to about $10\\times 2\\times 2\n{\\rm mm}^3$. We determined, via single crystal\nx-ray-diffraction-structure refinement and electron-microprobe\nanalysis (employing single crystals of GdFe$_2$ as standards), the\ncrystals to be slightly iron deficient, but detected no Al on Fe\nsites or vice versa (measured compositions were between\nGdFe$_{3.88(5)}$Al$_8$ and GdFe$_{3.96(1)}$Al$_8$; deviations of\nthe Al stoichiometry from $8$ were always less than 2 standard\ndeviations and are not listed).\n\nWe note that the x-ray-diffraction-structure refinement cannot\ndistinguish between a Fe deficiency due to vacancies and a\n(larger) Fe deficiency due to Al on Fe sites, both scenarios can\ngive an equal electron density. However, assuming an Al\/Fe mixture\nto retain the full occupancy of the Fe site leads (for one of the\ninvestigated crystals) to a GdFe$_{3.92(1)}$Al$_{8.08(1)}$\ncomposition and, thus, to an excess of Al. Since the\nelectron-microprobe analysis results indicated no significant\nvariations in the $1\/8$ ratio of Gd to Al, we did the structure\nrefinements under the assumption that the electron density on the\nFe sites is caused by vacancies, although some Al\/Fe substitution\ncould never be excluded. Whatever scenario is chosen, none of our\nconclusions drawn in this article are affected by the issue of\nwhether or not some Al atoms are present on the Fe site, and for\ndefiniteness subsequently we will assume that the vacancy scenario\nis the correct one. The composition of our samples can thus be\ndescribed with the empirical formula GdFe$_{4-\\delta}$Al$_8$, with\n$\\delta \\approx 0.04-0.12$. Surface scans in the\nelectron-microprobe analysis gave no indications of compositional\nvariations within the same crystal.\n\nSince the exact iron stoichiometry varies slightly from crystal to\ncrystal (even within the same growth batch) and this was found to\nhave a significant influence on the magnetic properties (see Sec.\\\n\\ref{lowH} below) most of the measurements presented below were\nperformed on the same crystal, sample I, which has a refined\ncomposition GdFe$_{3.96(1)}$Al$_8$ ($a=8.7699(9)\\, $\\AA,\n$c=5.0440(6)\\, $\\AA). A bar cut from the crystal by a wire saw was\nconnected with contacts for 4-point electrical-transport\nmeasurements with the current density $j\\|$[110] (sample Ia).\nOther samples were connected with contacts for electrical\ntransport measurements with either $j\\|$[110] or $j\\|$[001]. On a\nsecond bar (sample Ib, which had a mass of $2.22\\,{\\rm mg}$) cut\nfrom crystal I we measured the longitudinal magnetization and the\nzero-field specific heat. In order to investigate the relation\nbetween compositional variations and variations in the transition\ntemperatures, we carried out additional magnetization measurements\non several other crystals, the composition of which was determined\nby x-ray diffraction and electron microprobe. The additional\ncrystals are labelled II to VII in the order of their appearance\nin the text. The above measurements were performed with commercial\n(Quantum Design) laboratory equipment. For the high-field\nmeasurements, the sample orientation was fixed with a\ntwo-component glue (X60 from {W{\\\"{a}}getechnik} GmbH, Darmstadt,\nGermany).\n\nSample Ib was also used in the XRMS experiment, performed on the\n6ID-B undulator beamline in the MUCAT sector at the Advanced\nPhoton Source, Argonne National Laboratory. The incident energy of\nx-rays was tuned to the Gd $L_{\\rm II}$ edge ($E=7.934\\,{\\rm\nkeV}$), using a liquid nitrogen cooled, double crystal Si (111)\nmonochromator and a bent mirror. Sample Ib was mounted on a copper\nrod on the cold finger of a closed cycle displex refrigerator,\nsuch that a natural (110) facet was exposed to the x-ray beam.\nThermal transfer was enhanced by embedding the sample in copper\npaste. The sample, oriented so that the scattering plane of the\nexperiment was coincident with the $a-b$ plane, was encapsulated\nin a Be dome with He exchange gas to further enhance thermal\ntransfer. The incident radiation was linearly polarized\nperpendicular to the scattering plane ($\\sigma$ polarized). In\nthis geometry, only the component of the magnetic moment that is\nin the scattering plane contributes to the resonant magnetic\nscattering arising from electric dipole transitions ($E1$) from\n$2p$ to $5d$ states. Furthermore, the dipole resonant magnetic\nscattering rotates the linear polarization into the scattering\nplane ($\\pi$ polarization). In contrast, charge scattering does\nnot change the polarization of the incident beam\n($\\sigma\\rightarrow\\sigma$ scattering). Pyrolytic graphite PG\n(006) was used as a polarization analyzer, selecting primarily\n$\\pi$ polarized radiation. For $E=7.934\\,{\\rm keV}$, the\npolarization analyzer used reduces the detected intensity\nresulting from $\\sigma\\rightarrow\\sigma$ charge scattering by\nabout $99.9\\%$, whereas the $\\sigma\\rightarrow\\pi$ resonant\nmagnetic scattering is passed with little loss. Thus, the\npolarization analysis suppresses the charge scattering relative to\nthe magnetic scattering signal.\n\n\n\\label{res}\n\\section{Low field measurements and influence of Fe content}\n\\label{lowH}\n\\begin{figure}[tb]\n\\includegraphics[width=0.98\\linewidth]{xlHTP.eps}\n\\caption{(Color online) Zero-field resistivity $\\rho$ (along\n[110], sample Ia) and inverse dc susceptibility $\\chi ^{-1}$ vs\ntemperature $T$ (sample Ib). } \\label{Fig2}\n\\end{figure}\n\nIn this section, we will first present zero-field transport and\nlow field magnetization measurements on our main crystal (samples\nIa\/Ib). Then, we will briefly discuss sample-to-sample differences\nin the characteristic temperatures, which we think are related to\nFe deficiency. Finally, we will present specificheat measurements\non sample Ib indicating that the Gd $4f$ moments are ordered only\nat low temperatures. The main implications of the measurements\nwill be discussed later in Sec.\\ \\ref{disc}.\n\nIn Fig.\\ \\ref{Fig2}, we show, for sample Ia, the $T$ dependence of\nthe resistivity $\\rho$ ($j\\|$[110]). Since the sample is rather\nsmall, the estimated uncertainty in the absolute value of the\nresistivity is about 15\\%. The antiferromagnetic ordering at $T_N\n\\simeq 155\\, {\\rm K}$ is visible as a sharp kink in $\\rho\n(T)$.\\cite{note_Palasyuk04} The inverse dc magnetic susceptibility\n$\\chi^{-1}$ ($M\/H$) of sample Ib, cut from the same crystal as\nsample Ia, is also presented in Fig.\\ \\ref{Fig2}. The dashed and\ndotted curves shown are determined from the magnetization $M(T)$\nin a field of $\\mu_0 \\,H=0.1\\,{\\rm T}$ applied parallel to [110]\nand parallel to [001]. Below this field $M$ vs $H$ is linear for\n$T \\gtrsim 40\\, {\\rm K}$. Since $\\chi$ is slightly anisotropic, we\nalso calculated the polycrystalline average $\\chi_{\\rm\npoly}=(2\\chi_{110}+\\chi_{001})\/3$, shown as squares. In\n$\\chi^{-1}$ for both directions the antiferromagnetic ordering is\nmanifest by a change in slope, but it is much less sharp than the\ncorresponding signature in the resistivity. A weak signature of\nthe {N{\\'{e}}el} transition may be expected if, as we will discuss\nbelow, only the Fe moments order at $T_N$. In agreement with\nmeasurements on polycrystalline samples,\\cite{Buschow78} $\\chi$\nfollows Curie-Weiss behavior both above and, over a limited $T$\nrange, below $T_N$. We found (on the calculated polycrystalline\naverage) above $T_N$ a Weiss temperature of $-165\\,{\\rm K}$ and an\neffective moment of $11.4 \\, \\mu_B \\, \/ {\\rm f.u.}$. Assuming\ncontributions by the magnetic atoms that are additive in the Curie\nconstant and a contribution by Gd of $7.9 \\, \\mu_B \\, \/ {\\rm Gd}$,\nthe free moment per Fe atom in the paramagnetic state is $4.1 \\,\n\\mu_B \\, \/ {\\rm Fe}$. Below $T_N$ (range $70-115\\,{\\rm K}$) we\nfound a Weiss temperature of $17\\,{\\rm K}$ and an effective moment\nof $7.4 \\, \\mu_B \\, \/ {\\rm f.u.}$. The latter is closer to the Gd\nfree-ion value ($7.9 \\, \\mu_B \\, \/ {\\rm f.u.}$) than the one\nreported by Buschow {\\em et al.}\\cite{Buschow78} ($6.2 \\, \\mu_B \\,\n\/ {\\rm f.u.}$).\n\n\\begin{figure}[t!b]\n\\includegraphics[width=0.98\\linewidth]{xlRcomps.eps}\n\\caption{(Color online) Low-temperature normalized resistance\n(panel a) and derivative of high temperature resistance (panel b)\nvs temperature $T$ of various crystals: Ia (thick line) and II\n(dotted line) for $j\\|$[110], III (open circles) and IV (full\ncircles) for $j\\|$[001].} \\label{Fig3}\n\\end{figure}\n\nThe resistivity curve shows additional features just below $30\\,\n{\\rm K}$, magnified in Fig.\\ \\ref{Fig3}a) (thick line): The broad\npeak in $\\rho (T)$ at $\\sim \\! \\! 28\\, {\\rm K}$ with a drop at\n$\\sim \\! \\! 26.5\\,{\\rm K}$ ($T_2$) indicates a phase transition.\nFurthermore, with decreasing $T$, at $\\sim \\! \\! 21\\, {\\rm K}$\n($T_1$) there is a sudden increase in $\\rho$ by $\\sim 10\\%$. This\nfeature is hysteretic in temperature, suggesting that the feature\nis associated with a first-order transition.\\cite{note_Fujiwara}\nNote that the hysteresis visible at $T_2$ is opposite to what is\nexpected for a first-order transition (i.e.,\nsuperheating\/supercooling). It is most likely a remnant of the\nhysteresis associated with the $T_1$ transition. At $T_1$, the\nform of the hysteresis is consistent with\nsuperheating\/supercooling.\n\nAlso shown in Fig.\\ \\ref{Fig3}a) are low-$T$ zero-field\nresistivity curves for three additional samples, one also with\n$j\\|$[110] and two with $j\\|$[001]. Whereas the feature of a\nsudden increase of $\\rho$ with decreasing $T$ is visible in all\ncurves, the other feature at slightly higher temperatures is only\nidentifiable when $j\\|$[110]. Based on the results from sample Ia\nonly, it may be tempting to associate the increase in $\\rho$ (with\ndecreasing $T$) with the opening of a superzone\ngap.\\cite{Mackintosh62} However, the presence of this increase\nwith the same order of magnitude for both $j\\|$[110] and\n$j\\|$[001] makes a superzone gap scenario less likely.\n\nIt is also clear from Fig.\\ \\ref{Fig3}a) that there are\nsignificant variations of the temperatures $T_1$ corresponding to\nthis feature. For comparison, Fig.\\ \\ref{Fig3}b) shows the\nresistivity derivative ${\\rm d} \\rho \/ {\\rm d} T$ for the same\nsamples at higher $T$, in order to make the {N{\\'{e}}el}\ntransition more visible, which manifests itself differently for\nthe two current density directions. The curves indicate a sharp,\nwell defined $T_N$ for each sample and comparing Figs.\\\n\\ref{Fig3}a) and \\ref{Fig3}b) suggests a correlation between\nhigher $T_1$ and lower $T_N$.\n\n\n\\begin{figure}[tb]\n\\includegraphics[width=0.98\\linewidth]{xlMTmicroprobe.eps}\n\\caption{(Color online) Low field ($H\\|[001]$) Magnetization of\nvarious crystals with Fe stoichiometry (as described in Sec.\\\n\\ref{exp}) determined by electron microprobe or\nx-ray-diffraction-structure refinement.} \\label{Fig4}\n\\end{figure}\n\nFigure \\ref{Fig4} displays low field ($H\\|$[001]) magnetization\nmeasurements performed on three samples with different Fe\nstoichiometry as determined by x-ray diffraction and electron\nmicroprobe: samples Ib (GdFe$_{3.96(1)}$Al$_8$, thick line), V\n(GdFe$_{3.94(6)}$Al$_8$, full squares), and VI\n(GdFe$_{3.88(5)}$Al$_8$, dashed line). The data were taken for\nboth decreasing and increasing $T$ (field cooled), and\nnormalized\\cite{note_normalizedM} to the maximum in magnetization.\nA comparison with Fig.\\ \\ref{Fig3}a) indicates that the rise in\n$M$ with $T$ is associated with $T_1$, the subsequent decrease\nwith $T_2$. A decrease in the Fe content seems to systematically\nshift $T_1$ and $T_2$ to higher temperatures. We note that for\ncrystal I $M(T)$ (Fig.\\ \\ref{Fig4}) indicates a higher $T_2$ than\n$\\rho (T)$ [Fig.\\ \\ref{Fig3}a)]. In principle, this might be due\nto a different Fe content of the two pieces (Ia and Ib) cut from\nthe crystal and used for these measurements. However, we do not\nbelieve that this is the case, because the specific-heat peak\nassociated with $T_2$ (see below; measured on Ib) corresponds well\nto $T_2$ deduced from $\\rho (T)$ on sample Ia.\n\nCombining the above data it seems that decreasing the Fe content\nresults in a decrease of $T_N$ and an increase of $T_1$ and $T_2$.\nIn order to confirm this, a study involving samples that are much\nmore Fe deficient, would be highly desirable.\n\n\\begin{figure}[tb]\n\\includegraphics[width=0.98\\linewidth]{nxlHCalt.eps}\n\\caption{(Color online) %\nSpecific heat $C_p$ measured on sample Ib in zero field\n($\\blacksquare$) and Gd $4f$ moment contribution to the specific\nheat ($\\circ$), obtained by subtracting the specific heat of\nYFe$_4$Al$_8$ (full line, from Ref.\\ \\onlinecite{Hagmusa98}).\nInset: Entropy of the Gd $4f$ moments obtained by integrating the\n$4f$ moment contribution to $C_p \/ T$. The two lines shown were\ncalculated using two extrapolations of the specific heat to $T=0$,\nindicated by dashed lines in the main panel (see text for details\nof the extrapolations). The dotted line in the inset indicates the\nfull Gd $4f$ entropy.} \\label{Fig5}\n\\end{figure}\n\n\nFigure \\ref{Fig5} shows the measured specific heat $C_p$ of sample\nIb (closed squares) after subtraction of the addenda contribution\n(the specific heat of sample platform and grease measured before\nmounting the sample). Also shown is the mass-scaled specific heat\nof YFe$_4$Al$_8$ from Ref.\\ \\onlinecite{Hagmusa98} (full line),\nand the difference between the two specific heats (open circles).\nSince the necessary mass-scaling is small, the remaining specific\nheat is close to the magnetic contribution associated with the Gd\nsublattice. The broad, asymmetric, peak at $\\sim \\! \\! 26.5 \\,\n{\\rm K}$ corresponds to the drop in $\\rho$ at $T_2$ [determined on\nsample Ia cut from the same crystal as sample Ib, see Fig.\\\n\\ref{Fig3}a) dashed curve]. The broadness of the specific-heat\npeak may suggest a degree of inhomogeneity of the Fe distribution\nin the sample, although the electron-microprobe analysis provided\nno indications of inhomogeneities within crystals.\n\n\nNo feature in the specific heat is visible around $T_1$, the small\ndiscontinuity at $20\\,{\\rm K}$ indicated by the data was found to\nbe an artifact due to the change of the heat pulse intensity\napplied by the system. We checked the specific-heat raw data, and\ncan exclude any latent heat restricted to a temperature region\nsmaller than the spacing of measurement points. We conclude that\nthe latent heat of the $T_1$ transition has to be small, i.e., the\nphases above and below $T_1$ have similar entropies. The\nalternative explanation of the specific heat, that $T_1$ is not a\nfirst-order transition, seems unlikely in light of not only the\npresence, but also the shape of thermal hysteresis observed in\n$\\rho (T)$, which is particularly clear when a magnetic field is\napplied in the [110] direction (see Sec.\\ \\ref{highH} below).\n\nThe corresponding magnetic entropy $S$, obtained by integration,\nis shown in the inset of Fig.\\ \\ref{Fig5}. The two curves shown\nare calculated with two different extrapolations of $C_p$ to zero\ntemperature (dashed lines in the main panel). One extrapolation is\nobtained by connecting the $C_p$ data point at the lowest $T$\nlinearly to $C_p=0$ at zero temperature, the other by connecting\nthe $C_p\/T$ data at the lowest $T$ linearly to $C_p\/T=0$ at zero\ntemperature. Given typically observed specific heats of Gd\ncompounds,\\cite{note_DuongGdFe2Ge2} the actual specific heat at\nlow temperature will most likely be between these two\nextrapolations, and thus the two curves shown in the inset of\nFig.\\ \\ref{Fig5} may be considered lower and upper limits of the\nGd $4f$ entropy (neglecting additional experimental\nuncertainties). At $30\\, {\\rm K}$ the magnetic entropy already\nreaches 80\\% (78\\% and 86\\% for the two low-$T$ extrapolations\nmade) of the full entropy of Gd $4f$ moments (${\\rm R}\\,\\ln 8$,\ndashed line). Above $45\\, {\\rm K}$ it hardly varies anymore,\nhaving reached $>90\\%$ and $>97\\%$ of the full $4f$ entropy for\nthe two low-$T$ extrapolations made. This strongly suggests that\nthe Gd $4f$ moments are not ordered above $T_2$, and the\ntransition at $T_2$ corresponds to the ordering of the $4f$\nmoments.\n\nBy careful measurement and analysis of the low-field thermodynamic\nand transport data of solution grown single crystals of Gd148 we\nhave shown that there are sharp features associated with two\nmagnetic phase transitions. These transitions appear to be\nprimarily related to the ordering of the Gd sublattice. In\naddition we have been able to establish a clear correlation\nbetween small variations in the Fe stoichiometry and $T_N$, $T_1$,\nand $T_2$. Having done this, we will now focus on the sample that\nis closest to full stoichiometry (crystal I) and try to more fully\ndelineate and understand the field and temperature dependence and\nthe structure of the magnetic phases in this sample.\n\n\n\\section{High-field measurements}\n\\label{highH}\n\nApplying a magnetic field often helps to clarify the nature of\nmagnetic phases observed in zero (or low) field measurements. We\ntherefore measured the field dependent electrical resistivity and\nthe magnetization on several samples. Shown in this section are\nelectrical transport (subsection \\ref{highHrho}) and magnetization\n(subsection \\ref{highHM}) data measured on samples Ia\/Ib at low\ntemperatures in fields applied in-plane and out-of-plane. The\nfeatures in resistivity corresponding to the two low-temperature\ntransitions are much sharper when an in-plane field is applied and\nobserved features suggest complex domain effects. The\nmagnetization data are used to estimate the in-plane and\nout-of-plane components of the spontaneous ferromagnetic moment.\nThe phase diagrams for $H\\|$[110] and $H\\|$[001], which can be\nconstructed from these transport and magnetization data, will be\npresented and discussed in Sec.\\ \\ref{disc}.\n\n\\subsection{Electrical transport}\n\\label{highHrho}\n\\begin{figure}[tb]\n\\includegraphics[width=0.98\\linewidth]{xldomains.eps}\n\\caption{(Color online) resistivity $\\rho$ vs $T$ in fields of\n$3.5\\,{\\rm T}$ applied in-plane parallel or perpendicular to the\ncurrent density $j\\|[110]$ (sample Ia). Arrows indicate the $T$\ndirection in which the measurements were performed. The open\ncircle indicates the value of the $T_1$ transition in $3.5\\,{\\rm\nT}$, averaged for $T$ increasing and decreasing.} \\label{Fig6}\n\\end{figure}\n\nFigure \\ref{Fig6} shows the resistivity $\\rho$ ($j\\|$[110]) of\nsample Ia when a field of $3.5\\,{\\rm T}$ is applied either $\\|j$\n(longitudinal resistivity) or $\\|[1\\overline{1}0]$ (i.e.\\\ntransverse resistivity, $\\perp$$j$, but in a crystallographically\nequivalent in-plane direction). For both field directions, the\nmagnitude of the jump at $T_1$ is enhanced and the hysteresis in\ntemperature greatly increased, making the first-order nature of\nthe transition more apparent. The average\n($T\\uparrow$,$T\\downarrow$) temperature of the $T_1$ feature,\nhowever, is only weakly influenced by an in-plane field. In\ncontrast to this, the $T_2$ feature is shifted to higher $T$ by\n$H\\|$[110], suggesting a ferromagnetic nature of the phase at\n$T0$. Replacing the variable $k_T^{'} \\rightarrow k_T^{'} +k_T$ one can write\n\\begin{equation}\n\t\\label{14}\n\t\\int\\frac{\\text{d}^2k_T^{'}}{(k_T^{'}-k_T)^2}\\frac{1}{(k_T^{'}-k_T)^2+k_T^{'2}}=\\int\\frac{\\text{d}^2k_T^{'}}{k_T^{'2}}\\frac{1}{k_T^{'2}+(k_T^{'}+k_T)^2}.\n\\end{equation}\nNow using \\eqref{13} in the r.h.s. of \\eqref{14} we get\n\\begin{equation}\n\t\\label{15}\n\t\\begin{split}\n\t\t\\int\\frac{\\text{d}k_T^{'} k_T^{'} \\text{d}\\theta}{k_T^{'2}(2k_T^{'2}+k_T^2+2k_T^{'}k_T\\cos\\theta)}&=2\\pi\\int\\frac{\\text{d}k_T^{'} k_T^{'}}{k_T^{'2} \\sqrt{4k_T^{'4}+k_T^4}}\\\\&=\\pi\\int\\frac{\\text{d}k_T^{'2}}{k_T^{'2} \\sqrt{4k_T^{'4}+k_T^4}}.\n\t\\end{split}\n\\end{equation}\nSimilarly using \\eqref{13} in the l.h.s. of \\eqref{14} we get\n\n\\begin{equation}\n\t\\label{16}\n\t\\begin{split}\n\t\t\\int\\frac{\\text{d}^2k_T^{'}}{(k_T^{'}-k_T)^2}\\frac{1}{(k_T^{'}-k_T)^2+k_T^{'2}} &=\\pi \\int \\frac{dk_T^{'2}}{k_T^{'2}}\\frac{1}{\\left| k_T^{'2}-k_T^2\\right| }\\\\& -\\pi\\int\\frac{\\text{d}k_T^{'2}}{k_T^{'2} \\sqrt{4k_T^{'4}+k_T^4}}.\n\t\\end{split}\n\\end{equation}\nEquating \\eqref{15} and \\eqref{16} we obtain\n\n\\begin{equation}\n\t\\label{17}\n\t\\int\\frac{dk_T^{'2}}{k_T^{'2}}\\frac{1}{\\left| k_T^{'2}-k_T^2\\right| }= 2 \\int\\frac{\\text{d}k_T^{'2} }{k_T^{'2} \\sqrt{4k_T^{'4}+k_T^4}},\n\\end{equation}\nwhich turns out to be well-behaved for both integration limit for $I_1$. On the other hand, in the limit $k_{T_{\\text{min}}}^{'2} \\leq k_T^{'2}\\leq k_T^2$ i.e. for not too large $k_T^{'2}$, the contribution from the longitudinal component to the gluon virtuality $k^{'2}$ becomes negligible which in turn preserves the no ordering condition of transverse momentum of BFKL kinematics i.e. $k_T^{'2}\\approx k_T^2$ very strictly. Therefore, in this limit $k_{T_{\\text{min}}}^{'2} \\leq k_T^{'2}\\leq k_T^2$, it is justified to implant a factor $\\left(k_T^{2}\/k_T^{'2}\\right)^\\lambda$ inside the integrals which will in fact make our calculation simpler without altering the underlying physics. Moreover, this will fix the infrared divergence problem of the second improper integral $I_2$ by allowing us to evaluate the integral down to $k_{T_\\text{min}}^{'2}$.\n\\par Now considering these approximations in \\eqref{11} and \\eqref{12} and putting $I_1$ in \\eqref{12} we obtain\n\\begin{align}\n\t\\label{18}\n\t\\begin{split}\n\t\t\\xi(k_T^2)=&\\frac{\\alpha _s N_c k_T^2}{\\pi }\\bigg[\\int _{k_{T_{\\min }}^{'2}}^{k_T^2}\\frac{dk_T^{'2}}{k_T^{'2}}\\left(\\frac{k_T^2}{k_T^{'2}}\\right)^\\lambda\\frac{k_T^{'2}-k_T^2}{\\left| k_T^{'2}-k_T^2\\right| }\\\\&+\\int _{k_T^{2}}^{\\infty}\\frac{dk_T^{'2}}{k_T^{'2}}\\left(\\frac{k_T^2}{k_T^{'2}}\\right)^{\\lambda }\\frac{k_T^{'2}-k_T^2}{\\left| k_T^{'2}-k_T^2\\right| }\\bigg]\\\\=&\\frac{\\alpha_s N_c k_T^2}{\\pi}\\frac{1}{\\lambda}\\left(2-k_T^{2\\lambda}\\right)\\approx-\\frac{\\alpha_s N_c}{\\pi\\lambda}(k_T^2)^{\\lambda+1},\n\t\\end{split}\n\\end{align}\n\n\\begin{align}\n\t\\label{19}\n\t\\begin{split}\n\t\t\\zeta(k_T^2)=&\\frac{\\alpha _s N_c }{\\pi }k_T^2\\bigg[\\int _{k_{T_\\text{min}}^{'2}}^{\\infty }\\left(\\frac{k_T^2}{k_T^{'2}}\\right)^\\lambda\\frac{dk_T^{'2}}{k_T^{'2}}\\frac{1}{\\left| k_T^{'2}-k_T^2\\right| }\n\t\t\\\\&-\\int _{k_{T_{\\min }}^{'2}}^{\\infty }\\frac{dk_T^{'2}}{k_T^{'2}}\\frac{1}{\\sqrt{k_T^4+4k_T^{'4}}}\\bigg]\\\\\n\t\t=&\\frac{\\alpha _s N_c }{\\pi }\\bigg[\\frac{k_T^{2\\lambda}}{\\lambda}- 2^{2-\\frac{\\lambda }{2}} \\lambda^{-1}k_T^{2\\lambda}\\bigg(1-\\frac{\\sqrt{k_T^4+4}}{k_T^2}\\bigg)^{\\lambda \/2} \\, _2F_1\\\\&-\\lambda\\ln \\left(\\frac{k_T^2}{2}+\\sqrt{1+\\frac{k_T^4}{4}}\\right)\\bigg]\\approx \\frac{\\alpha _s N_c }{\\pi }\\left(\\epsilon+\\frac{(k_T^2)^\\lambda }{\\lambda}\\right),\n\t\\end{split}\n\\end{align}\nwhere $\\, _2F_1$=$\\, _2\\text{F}_1\\left(-\\frac{\\lambda }{2},\\frac{\\lambda }{2};1-\\frac{\\lambda }{2};\\frac{k_T^2+\\sqrt{k_T^4+4}}{2 k_T^2}\\right)$ is a standard hypergeometric function. In the above calculations we have taken infrared cutoff $k_{T_\\text{min}}^{'2}= 1 \\text{ GeV}^2$ since for unified BFKL-DGLAP framework this provided a very consistent result towards HERA DIS data for proton structure function $F_2$ \\cite{14}. In \\eqref{19} $(k_T^2)^\\lambda\/\\lambda$ is the only dominant term since other terms are significantly small. The contribution from the logarithmic term in \\eqref{19} is negligible in comparison to the net contribution for all $k_T^2$. However, phenomenological studies shows the term involving hypergeometric function in \\eqref{19} becomes irrelevant towards change in $k_T^2$ i.e. it possess a constant value $(\\approx -4.79)$ (see Fig.~\\ref{reg}(a)). This constant contribution is insignificant to the net contribution for $\\zeta(k_T^2)$ if $k_T^2$ is enough high. But for small $k_T^2$ this contribution can not be neglected since at this range, the $k_T^{2\\lambda}\/\\lambda$ contribution itself is small. In consequence, this constant contribution can be treated as a small perturbation $\\epsilon$ to the dominant term $k_T^{2\\lambda}\/\\lambda$ . We have performed phenomenological determination of the constant perturbation parameter $\\epsilon$ using standard non-linear regression method (see Fig.~\\ref{reg}(b)).\n\n\\begin{figure*}[tbp]\n\t\\centering\n\t\\includegraphics[width=.38\\textwidth,clip]{epsilon.eps}\\hspace{1cm}\n\t\\includegraphics[width=.38\\textwidth,clip]{N11.eps}\n\t\\caption{\\label{reg} (a) $k_T^2$ dependence of the small perturbation $\\tilde{\\epsilon}=|\\epsilon|$ corresponding to the term with the hypergeometric function in \\eqref{9} (left). (b) A phenomenological calculation of $\\epsilon$ using standard non-linear regression (right). Comparison between $\\tilde{\\zeta}=\\frac{\\pi}{\\alpha_sN_c}\\zeta$ (green dotted line) and $\\epsilon+2\\lambda^{-1}k_T^{2\\lambda}$ (red dotted line) is shown.}\n\\end{figure*}\n\n\nNow we are set to solve our original equation \\eqref{10} which is indeed a 1st order semilinear (nonlinear) PDE. Our analytical approach of solving the same involves two steps: first we will express the nonlinear PDE in terms of a linear PDE then we will solve the linear PDE via. change of coordinates.\n\\par Substitution of $f(x, k_T^2)$ by its inverse function $\\omega(x,k_T^2)$ i.e. $f=\\omega^{-1}$ in \\eqref{10} yields\n\\begin{equation}\n\t\\label{20}\n\t-x\\frac{\\partial \\omega }{\\partial x}=\\xi \\frac{\\partial \\omega }{\\partial k_T^2}-\\zeta \\omega +\\Delta,\n\\end{equation}\nwhich is in fact a linear PDE in $\\frac{\\partial \\omega}{\\partial x}$, $\\frac{\\partial \\omega}{\\partial k_T^2}$\nand $\\omega$. Now we construct a new set of co-ordinate $\\sigma\\equiv\\sigma(x, k_T^2)$ and $\\eta\\equiv\\eta(x,k_T^2)$ such that it transforms \\eqref{20} into an ODE. To be specific we define this transformation in such a way that it is one to one for all $(x,k_T^2)$ in some set of points D in $x$-$k_T^2$ plane. This will allow us to solve \\eqref{20} for $x$ and $k_T^2$ as functions of $\\sigma$ and $\\eta$. The only requirement is that we should ensure the Jacobian of the transformation does not vanish i.e.\n$\\text{J}=\\left|\n\\begin{array}{cc}\n\\sigma_x & \\eta_x \\\\\n\\sigma_{k_T^2} & \\eta_{k_T^2} \n\\end{array} \\right| \\neq 0$ in D.\\\\\n\\par Next we want to recast \\eqref{20} in $(\\sigma,\\eta)$ plane computing the derivatives via chain rule:\n\\begin{equation}\n\t\\label{21}\n\t\\begin{split}\n\t\t&\\omega_x=\\omega_\\sigma\\sigma_x+\\omega_\\eta\\eta_x, \\\\\n\t\t&\\omega_{k_T^2}=\\omega_\\sigma\\sigma_{k_T^2}+\\omega_\\eta\\eta_{k_T^2}.\n\t\\end{split}\n\\end{equation}\nSubstitution of \\eqref{21} into \\eqref{20} yields\n\\begin{equation}\n\t\\label{22}\n\t-(x \\sigma_x+\\xi\\sigma_{k_T^2})\\omega_\\sigma-(x\\eta_x+\\xi\\eta_{k_T^2})\\omega_\\eta+\\zeta\\omega-\\Delta=0.\n\\end{equation}\nSince we want above equation to be expressed as an ODE, we require either,\n\\begin{equation*}\n\tx \\sigma_x+\\xi\\sigma_{k_T^2}=0 \\text{ or } x\\eta_x+\\xi\\eta_{k_T^2}=0.\n\\end{equation*}\nIf we consider $x \\sigma_x+\\xi\\sigma_{k_T^2}=0$, the solution $\\sigma$ is constant along the curves that satisfy\n\\begin{equation}\n\t\\label{23}\n\t\\frac{\\text{d}k_T^2}{\\text{d}x}=\\frac{\\xi}{x} \\implies \\ln (x C_1)=\\int\\frac{\\text{d} k_T^2}{\\xi}\n\t\\implies C_1=\\frac{e^{-\\frac{k_T^{-2\\lambda}}{l \\lambda}}}{x},\n\\end{equation}\nwhere $l=-\\frac{\\alpha_s N_c}{\\pi \\lambda}$ and $C_1$ is the constant of integration. Now we can choose the new coordinates as $\\sigma =\\frac{e^{-\\frac{k_T^{-2\\lambda}}{l \\lambda}}}{x}$ and $\\eta=x$. Here $\\text{J}=\\left|\n\\begin{array}{cc}\n\\sigma_x & \\eta_x \\\\\n\\sigma_{k_T^2} & \\eta_{k_T^2} \n\\end{array} \\right| \\neq 0$ as required.\n\\par We rewrite \\eqref{21} as\n\\begin{align}\n\t\\label{24}\n\t\\begin{split}\n\t\t&\\omega_x=\\omega_\\sigma\\sigma_x+\\omega_\\eta\\eta_x=-\\frac{e^{-\\frac{k_T^{-2\\lambda}}{l \\lambda}}}{x^2}\\omega_\\sigma+\\omega_\\eta \\text{,}\\\\\n\t\t&\\omega_{k_T^2}=\\omega_\\sigma\\sigma_{k_T^2}+\\omega_\\eta\\eta_{k_T^2}=\\frac{(k_T^2)^{-1-\\lambda}e^{-\\frac{k_T^{-2\\lambda}}{l \\lambda}}}{lx}.\n\t\\end{split}\n\\end{align}\nNow putting \\eqref{24} in \\eqref{22} we get\n\\begin{align}\n\t\\label{25}\n\t-\\eta \\omega_\\eta+\\zeta\\omega-\\Delta=0.\n\\end{align}\nEquation \\eqref{25} is an ODE and it can be solved using standard ODE solving techniques. Now solving \\eqref{25} and then transforming it to the original coordinates $(\\sigma,\\eta)\\rightarrow (x,k_T^2)$ we get the general solution of the KC improved MD-BFKL equation, \n\\begin{widetext}\n\t\\begin{align}\n\t\\begin{split}\n\t\\label{26}\n\t&f(x,k_T^2)=\\omega^{-1}(x,k_T^2)=\\frac{k_T^{-2\\frac{n}{l}} (-1)^{\\frac{n}{\\lambda l}} l^{-\\frac{n}{\\lambda l}} \\lambda ^{-\\frac{n}{\\lambda l}}}{\\text{G}\\left(\\frac{e^{-\\frac{k_T^{-2\\lambda}}{l \\lambda}}}{x}\\right) x^m+\\frac{18\\alpha_s ^2 (-1)^{1\/\\lambda } \\lambda ^{1\/\\lambda } N_c^2 l^{1\/\\lambda } e^{-\\frac{m k_T^{-2\\lambda }}{\\lambda l}} m^{-\\frac{\\lambda l+l+n}{\\lambda l}} \\Gamma \\left(\\frac{l+n}{l \\lambda }+1,-\\frac{k_T^{-2\\lambda } m}{l \\lambda }\\right)}{\\pi R^2 \\left(N_c^2-1\\right)}},\n\t\\end{split}\n\t\\end{align}\n\\end{widetext}\nwhere $\\Gamma \\left(\\frac{l+n}{l \\lambda }+1,-\\frac{k_T^{-2\\lambda } m}{l \\lambda }\\right)$ is a standard Gamma function and $\\text{G}\\left(\\frac{e^{-\\frac{k^{-2\\lambda}}{l \\lambda}}}{x}\\right)$ is an arbitrary continuously differentiable function. The parameters $m=\\frac{\\epsilon\\alpha_sN_c }{\\pi}$ and $n=\\frac{\\alpha_sN_c}{\\pi\\lambda}$ are coming from \\eqref{19}. In the following section, we try to get particular solutions for the PDE applying some initial boundary condition and present an analysis of the phenomenological aspects of \\eqref{26}.\n\n\n\n\\subsubsection{$x$ and $k_T^2$ evolution}\n\nIn this section, we study the small-$x$ dependence of gluon distribution by picking an appropriate input distribution at some high $x$, as well as the $k_T^2$ dependence of gluon distribution setting input distribution at some low $k_T^2$.\n\\par Let us rewrite \\eqref{26} rearranging a bit \n\\begin{widetext}\n\\begin{align}\n\t\\label{27}\n\t\\begin{split}\n\t\t\\text{G}\\left(\\frac{e^{-\\frac{k_T^{-2\\lambda}}{l \\lambda}}}{x}\\right)=\\frac{1}{N_c^2-1}\\Bigg[&x^{-m}N_C^2 \\bigg(\\frac{18\\alpha_s ^2 (-1)^{\\frac{1}{\\lambda }+1} \\lambda ^{1\/\\lambda } l^{1\/\\lambda } e^{-\\frac{m k_T^{-2\\lambda }}{\\lambda l}} m^{-\\frac{\\lambda l+l+n}{\\lambda l}} \\Gamma \\left(\\frac{l+n}{l \\lambda }+1,-\\frac{k_T^{-2\\lambda } m}{l \\lambda }\\right)}{\\pi R^2}\\\\\n\t\t&+\\frac{1}{f(x,k_T^2)} k_T^{-2\\frac{n}{l}} (-1)^{\\frac{n}{\\lambda l}} l^{-\\frac{n}{\\lambda l}} \\lambda ^{-\\frac{n}{\\lambda l}}\\bigg)+\\frac{1}{f(x,k_T^2)}x^{-m} k_T^{-2\\frac{n}{l}} (-1)^{\\frac{n}{\\lambda l}+1} l^{-\\frac{n}{\\lambda l}} \\lambda ^{-\\frac{n}{\\lambda l}}\\Bigg],\n\t\\end{split}\n\\end{align}\n\\end{widetext}First we will try to evaluate the functional form of the arbitrary differentiable function G applying some initial boundary condition on it. For $k_T^2$ evolution we set the initial distribution at $(x,k_0^2)$ where $x$ is fixed throughout the evolution. At $(x,k_0^2)$ let the argument of G be denoted by\n$t=\\frac{e^{-\\frac{k_0^{-2\\lambda}}{l \\lambda}}}{x}$.\n\nNow writing \\eqref{27} for $(x,k_0^2)$ and putting $x=\\frac{e^{-\\frac{k_0^{-2\\lambda}}{l \\lambda}}}{t}$ we get\n\\begin{widetext}\n\\begin{equation}\n\t\\label{28}\n\t\\begin{split}\n\t\t\\text{G}\\left(t\\right)=\\frac{1}{N_c^2-1}\\Bigg[&\\frac{e^{\\frac{mk_0^{-2\\lambda}}{l \\lambda}}}{t^{-m}}N_c^2 \\bigg(\\frac{18\\alpha_s ^2 (-1)^{\\frac{1}{\\lambda }+1} \\lambda ^{1\/\\lambda } l^{1\/\\lambda } e^{-\\frac{m k_0^{-2\\lambda }}{\\lambda l}} m^{-\\frac{\\lambda l+l+n}{\\lambda l}} \\Gamma \\left(\\frac{l+n}{l \\lambda }+1,-\\frac{k_0^{-2\\lambda } m}{l \\lambda }\\right)}{\\pi R^2}\\\\\n\t\t&+\\frac{1}{F_0} k_0^{-2\\frac{n}{l}} (-1)^{\\frac{n}{\\lambda l}} l^{-\\frac{n}{\\lambda l}} \\lambda ^{-\\frac{n}{\\lambda l}}\\bigg)+\\frac{1}{F_0}\\frac{e^{\\frac{mk_0^{-2\\lambda}}{l \\lambda}}}{t^{-m}} k_0^{-2\\frac{n}{l}} (-1)^{\\frac{n}{\\lambda l}+1} l^{-\\frac{n}{\\lambda l}} \\lambda ^{-\\frac{n}{\\lambda l}}\\Bigg],\n\t\\end{split}\n\\end{equation}\n\\end{widetext}\nwhere $F_0$ is the unintegrated gluon distribution at $(x,k_0^2)$.\nEquation \\eqref{28} is the functional form of G. Replacing $t$ with any other argument will give us the value of G at that particular argument. We put $t=\\frac{e^{-\\frac{k_T^{-2\\lambda}}{l \\lambda}}}{x}$ in \\eqref{28} which gives us the l.h.s. of \\eqref{27} and then we solve \\eqref{27} for $f(x,k_T^2)$. The solution for $k_T^2$ evolution turns out to be\n\\begin{widetext}\n\\begin{equation}\n\t\\label{29}\n\t\\begin{split}\n\t\tf(x,k_T^2)=\\dfrac{(-1)^{\\frac{n}{\\lambda l}} \\big(N_c^2-1\\big) k_T^{-2\\frac{n}{l}} k_0^{2n\/l} \\big( e^{\\frac{k_T^{-2\\lambda }-k_0^{-2\\lambda }}{\\lambda l}}\\big)^m\\text{A}[x,k_T^2]}{(-1)^{\\frac{1}{\\lambda }+1}N_c^2 \\text{B}[ k_0^2 ] +(-1)^{\\frac{n}{\\lambda l}}N_c^2 F_0 \\text{A}[x,k_T^2] +(-1)^{1\/\\lambda } N_c^2 \\text{B}[k_T^2]+ (-1)^{\\frac{n}{\\lambda l}+1}F_0 \\text{A}[x,k_T^2]},\n\t\\end{split}\n\\end{equation}\n\\end{widetext}\nwhere\n\\begin{align*}\n\t\\begin{split}\n\t\t&\\text{A}[x,k_T^2]=\\pi R^2 x^m m^{\\frac{\\lambda l+l+n}{\\lambda l}} e^{\\frac{m \\big(k_T^{-2\\lambda }+k_0^{-2\\lambda }\\big)}{\\lambda l}}; \\\\\n\t\t&\\text{B}[i]=18 \\alpha_s ^2 x^m l^{\\frac{l+n}{\\lambda l}} \\lambda ^{\\frac{l+n}{\\lambda l}} k_0^{2n\/l} \\Gamma\\left. \\left(\\frac{l+n}{l \\lambda }+1,-\\frac{m (i)^{-\\lambda }}{l \\lambda }\\right)\\right\\vert_{i=k_T^2 , k_0^2}.\n\t\\end{split}\n\\end{align*}\n\\begin{figure*}[t]\n\t\\centering\n\t\\includegraphics[trim=25 5 21.5 0,width=.46\\textwidth,clip]{x-4.eps}\\hspace{5mm}\n\t\\includegraphics[trim=25 5 21.5 0,width=.46\\textwidth,clip]{x-5.eps}\n\t\n\n\t\\caption{\\label{k1}$k_T^2$ evolution of unintegrated gluon distribution $f(x,k_T^2)$. Our results of KC improved MD-BFKL are shown for conventional $R=5 \\text{ GeV}^{-1}$ (gluons are distributed throughout the nucleus) and $R=2 \\text{ GeV}^{-1}$(at \"hot-spots\" within proton disk). Prediction from modified BK equation is plotted for comparison.}\n\\end{figure*}\n\\begin{figure*}[t]\n\t\\centering\n\t\\includegraphics[trim=25 5 21.5 0,width=.46\\textwidth,clip]{x-3.eps}\\hspace{5mm}\n\t\\includegraphics[trim=25 5 21.5 0,width=.46\\textwidth,clip]{x-6.eps}\n\t\n\n\t\\caption{\\label{k2}$Q^2$ evolution of collinear gluon distribution $xg(x,Q^2)$. Our results of KC improved MD-BFKL are shown for conventional $R=5 \\text{ GeV}^{-1}$ (gluons are distributed throughout the nucleus) and $R=2 \\text{ GeV}^{-1}$(at \"hot-spots\" within proton disk). Theoretical prediction is compared with global datasets NNPDF 3.1sx and CT 14.}\n\\end{figure*}\nSimilarly setting the initial distribution at $(x_0,k_T^2)$, we obtain the solution for x-evolution as follows\n\\begin{widetext}\n\\begin{equation}\n\t\\label{30}\n\tf(x,k_T^2)=\\dfrac{(-1)^{\\frac{3}{\\lambda }+\\frac{2 n}{\\lambda l}}\\left(N_c^2-1\\right) (\\frac{x_0}{x})^m k_T^{-2\\frac{n}{l}} \\left(-\\frac{k_T^{-2\\lambda }}{\\lambda l}-\\ln \\frac{x}{x_0}\\right)^{-\\frac{n}{\\lambda l} }\\text{C}[x,k_T^2]}\n\t{\\Bigg[\\splitdfrac{(-1)^{\\frac{3}{\\lambda }+\\frac{3 n}{\\lambda l}}N_c^2 F_0^{'} \\text{C}[x,k_T^2] +\\frac{ (-1)^{\\frac{2}{\\lambda }+\\frac{n}{\\lambda l}} }{k_T^2} \\Gamma_1 \\text{D}[x,k_T^2]}{+\n\t\t\t(-1)^{\\frac{2}{\\lambda }+\\frac{n}{\\lambda l}+1} k_T^{2\\left(\\frac{n}{l}-\\lambda \\left(\\frac{1}{\\lambda }+\\frac{n}{\\lambda l}\\right)\\right)} \\Gamma_2 \\text{D}[x,k_T^2]+(-1)^{\\frac{3}{\\lambda }+\\frac{3 n}{\\lambda l}+1}\\pi R^2 F_0^{'} x^m \\text{C}[x,k_T^2]}\\Bigg]},\n\\end{equation}\n\\end{widetext}\nwhere, $F_0^{'}$ is the initial gluon distribution at $(x_0,k_T^2)$,\n\\begin{equation*}\n\t\\begin{split}\n\t\t&\\Gamma_1=\\Gamma\\left(\\frac{l+n}{l \\lambda }+1,-\\frac{k_T^{-2\\lambda } m}{l \\lambda }\\right),\\text{ }\n\t\t\\\\&\\Gamma_2=\\Gamma\\left(\\frac{l+n}{l \\lambda }+1,m \\left(-\\frac{k_T^{-2\\lambda }}{l \\lambda }-\\ln \\frac{x}{x_0}\\right)\\right),\\\\\n\t\t&\\text{C}[x,k_T^2]=\\pi R^2 x^ml^{-\\frac{1}{\\lambda }-\\frac{n}{\\lambda l}} \\lambda ^{-\\frac{1}{\\lambda }-\\frac{n}{\\lambda l}} k_T^{2\\left(\\frac{n}{l}-\\lambda \\left(\\frac{1}{\\lambda }+\\frac{n}{\\lambda l}\\right)\\right)}\\\\\n\t\t&\\times m^{\\frac{2}{\\lambda }+\\frac{2 n}{\\lambda l}+1} \\left(-\\frac{k_T^{-2\\lambda }}{\\lambda l}-\\ln \\frac{x}{x_0}\\right)^{\\frac{1}{\\lambda }+\\frac{n}{\\lambda l}},\\\\\n\t\t&\\text{D}[x,k_T^2]=18 \\alpha_s ^2 N_c^2 x_0^m l^{-\\frac{n}{\\lambda l}} \\lambda ^{-\\frac{n}{\\lambda l}} e^{-\\frac{m k_T^{-2\\lambda }}{\\lambda l}}\\\\\n\t\t&\\times m^{\\frac{1}{\\lambda }+\\frac{n}{\\lambda l}} \\left(-\\frac{k_T^{-2\\lambda }}{\\lambda l}-\\ln \\frac{x}{x_0}\\right)^{1\/\\lambda }.\n\t\\end{split}\n\\end{equation*}\n\n\\begin{figure*}[t]\n\t\\centering\n\t\\includegraphics[trim=25 5 21.5 0,width=.46\\textwidth,clip]{k_5.eps}\\hspace{5mm}\n\t\\includegraphics[trim=25 5 21.5 0,width=.46\\textwidth,clip]{k_50.eps}\n\t\n\n\t\\caption{\\label{x1}$x$ evolution of unintegrated gluon distribution $f(x,k_T^2)$. Our results of KC improved MD-BFKL are shown for conventional $R=5 \\text{ GeV}^{-1}$ (gluons are distributed throughout the nucleus) and $R=2 \\text{ GeV}^{-1}$(at \"hot-spots\" within proton disk). Prediction from modified BK equation is plotted for comparison.}\n\\end{figure*}\n\\begin{figure*}[t]\n\t\\centering\n\t\\includegraphics[trim=25 5 21.5 0,width=.46\\textwidth,clip]{k_35.eps}\\hspace{5mm}\n\t\\includegraphics[trim=25 5 21.5 0,width=.46\\textwidth,clip]{k_100.eps}\n\t\n\n\t\\caption{\\label{x2}$x$ evolution of collinear gluon distribution $xg(x,k_T^2)$. Our results of KC improved MD-BFKL are shown for conventional $R=5 \\text{ GeV}^{-1}$ (gluons are distributed throughout the nucleus) and $R=2 \\text{ GeV}^{-1}$(at \"hot-spots\" within proton disk). Theoretical prediction is compared with global datasets NNPDF 3.1sx and CT 14.}\n\\end{figure*}\n\nWe have plotted the solution for both $x$ evolution \\eqref{30} and $k_T^2$ evolution \\eqref{29} in Fig.~\\ref{k1} and Fig.~\\ref{x1}. Our prediction of unintegrated gluon distribution $f(x,k_T^2)$ is contrasted with that of modified BK equation \\cite{5},\n\n\\begin{equation}\n\\label{bk}\n\\begin{split}\n&-x\\frac{\\partial f \\left(x, k_T^2\\right)}{\\partial x}=\\\\&\\frac{\\alpha_s N_c k_T^2}{\\pi}\\int_{k_{T_{\\min }}^{'2}}^{\\infty}\\frac{dk_T^{'2}}{k_T^{'2}}\\left[\\frac{f (x, k_T^{'2})-f \\left(x, k_T^2\\right)}{|k_T^{'2}-k_T^2|} + \\frac{f \\left(x, k_T^2\\right)}{\\sqrt{k_T^4+4k_T^{'4}}}\\right]\\\\\n&-\\alpha_{s}\\left(1-k_T^2\\frac{\\text{d}}{\\text{d}k_T^2}\\right)^2\\frac{k_T^2}{R^2}\\left[\\int_{k_T^2}^{\\infty}\\frac{\\text{d}k_T^{'2}}{k_T^{'2}}\\ln \\left(\\frac{k_T^{'2}}{k_T^2}\\right)f(x,k_T^2)\\right]\n\\end{split}\n\\end{equation}\nwhich is BFKL equation supplemented by the negative nonlinear term, derived in approximation of infinite and uniform target. In \\cite{5} the perturbative parton saturation is studied to a vast extent, including modification of \\eqref{bk} in terms of kinematic constraint, DGLAP, $\\text{P}_{gg}$ splitting function and running coupling constant, then solving the same numerically. We have also extracted collinear gluon distribution from unintegrated gluon distribution using the standard relation,\n\\begin{equation}\n\\label{f2g}\nxg(x,Q^2)=\\int_{0}^{Q^2}\\frac{\\text{d}k_T^2}{k_T^2}f(x,k_T^2)\n\\end{equation}\nsketched in Fig.~\\ref{k2} and Fig.~\\ref{x2}. Our predicted collinear gluon density is compared with that of LHAPDF global parameterization groups NNPDF 3.1sx \\cite{36} and CT 14 \\cite{53}. Both of the LHAPDF datasets include HERA as well as recent LHC data with high precision PDF sensitive measurements.\n\\begin{figure*}[t]\n\t\\centering\n\t\\includegraphics[trim=25 0 21.5 0,width=.47\\textwidth,clip]{compar_x.eps}\n\t\\hspace{1cm}\n\t\\includegraphics[trim=25 0 21 0,width=.46\\textwidth,clip]{compar_k.eps}\n\t\n\n\t\\caption{\\label{3a} (a) Comparison between linear BFKL, MD-BFKL without KC and MD-BFKL with KC for $k_T^2$ evolution at $x=10^{-4}$ and $10^{-6}$ (b) Comparison between linear BFKL, MD-BFKL without KC and MD-BFKL with KC for $x$ evolution at $k_T^2=5\\text{ GeV}^2$ and $50\\text{ GeV}^2$. Results are shown for $R= 5 \\text{ GeV}^{-1}$. }\n\\end{figure*} \n\\par Our prediction of both unintegrated and collinear gluon distribution is obtained for two different form of shadowing: conventional $R= 5 \\text{ GeV}^{-1}$ (order of proton radius) where gluons are spread throughout the nucleus and extreme $R= 2 \\text{ GeV}^{-1}$ where gluons are expected to concentrated in hotspots within the proton disk, recalling that $\\pi R^2$ is the transverse area within which gluons are concentrated inside proton. From Fig.~\\ref{k1}-\\ref{x2} it is clear that shadowing correction are more prominent when gluons are concentrated in hotspots within proton. \n\n\n\\par The $k_T^2$ (and $Q^2$) evolution in Fig.~\\ref{k1}-\\ref{k2} is studied for the kinematic range $1 \\text{ GeV}^2\\leq k_T^2(\\text{or }Q^2)\\leq 10^3 \\text{ GeV}^2$ corresponding to four different values of $x$ as indicated in the figure. Our evolution for both $f(x,k_T^2)$ and $xg(x,Q^2)$ shows a similar growth as modified BK as well as datasets respectively for all $x$. It is also observed that the growth of $f(x,k_T^2)$ (or $xg(x,Q^2)$) is almost linear for the entire kinematic range of $k_T^2$ (or $Q^2$). This is expected since the net shadowing term in \\eqref{8} has $1\/k_T^2$ dependence, which suppresses the contribution from the shadowing term at large $k_T^2$.\n\n\n\\par The $x$ evolution of $f(x,k_T^2)$ and $xg(x,Q^2)$ is shown in Fig.~\\ref{x1}-\\ref{x2} for two $k_T^2$ values viz. $5 \\text{ GeV}^2$, $50 \\text{ GeV}^2$ and two $Q^2$ values viz. $35 \\text{ GeV}^2$, $100 \\text{ GeV}^2$. The input is taken at higher $x$ value $x=10^{-2}$ and then evolved down to smaller $x$ value upto $x=10^{-6}$ thereby setting the kinematic range of evolution $10^{-6}\\leq x\\leq10^{-2}$. We observed that the singular $x^{-\\lambda}$ growth of the gluon is eventually suppressed by the net shadowing effect. KC improved MD-BFKL seem to poses a more intense shadowing then modified BK. However, it is hard to establish the existence of shadowing for $x\\geq 10^{-3}$. The obvious distinction between the two form of shadowing $R= \\text{5 GeV}^{-1}$ (conventional) and $R= 2\\text{ GeV}^{-1}$ (\"hotspot\") is also observed towards small-$x$ $(x\\leq 10^{-3})$. Interestingly towards very small-$x$ $(x\\leq 10^{-5})$, at small $k_T^2$ (or $Q^2$) values (viz. $k_T^2$=$5\\text{ GeV}^2$, $Q^2$=$35 \\text{ GeV}^2$) the gluon distribution becomes almost irrelevant of change in $x$. This could be a strong hint for the possible saturation phenomena in the small-$x$ high density regime.\n\n\\par In Fig.~\\ref{3a}(a) we have shown a comparison between the linear BFKL equation, MD-BFKL without kinematic constraint and MD-BFKL with kinematic constraint for $x$ evolution. The three solutions are compared for two different values of $k_T^2$ i.e. $ 5 \\text{ GeV}^2$ and $50 \\text{ GeV}^2$. Similarly in Fig.~\\ref{3a}(b) we have shown similar comparison but for $k_T^2$ evolution for two different values of $x$ i.e. $x=10^{-4}$ and $x=10^{-6}$. In Fig.~\\ref{3a} the singular $x^{-\\lambda}$ behavior of $f(x,k_T^2)$ is distinct for unshadowed linear BFKL equation. On the other hand, the deviation of the two different forms of the MD-BFKL equation from linear BFKL equation reflects the underlying shadowing correction. It is also observed that shadowing effect is more intense in MD-BFKL with KC than MD-BFKL without KC.\n\n\n\\subsubsection{Complete solution of KC improved MD-BFKL}\nIn this section we implant a functional form of the input distribution (more likely a dynamic one) on the general solution of our KC improved MD-BFKL equation \\eqref{26} and try to obtain a parametric form of the solution. The underlying motivation towards doing so is that this allows us to evolve our solution for both $x$ and $k_T^2$ simultaneously in $x\\text{-}k_T^2$ phase space which help us to portray a three-dimensional realization of the gluon evolution. \n\\par Recall the well-known solution of the linear BFKL equation \\cite{9}\n\\begin{equation}\n\t\\label{32}\n\tf(x,k_T^2)=\\beta \\frac{x^{-\\lambda}\\sqrt{k_T^2}}{\\sqrt{\\ln\\frac{1}{x}}}\\exp \\left(-\\frac{\\ln^2(k_T^2\/k_s^2)}{2\\Omega\\ln (1\/x)}\\right),\n\\end{equation}\nwhere $\\lambda= \\frac{3 \\alpha_s}{\\pi} 28 \\zeta(3)$, $\\zeta$ being Reimann zeta function and $\\Omega=32.1 \\alpha_s$. The nonperturbative parameter $k_s^2= 1 \\text{ GeV}^2$ and the normalization constant $\\beta\\sim 0.01$ \\cite{1}. Since far below saturation region both linear and nonlinear equation should give the same solution, therefore we can take \\eqref{32} as the input distribution for our general solution of KC improved MD-BFKL equation.\n\nFirst we try to find the functional form of the arbitrary differentiable function $\\text{G}\\left(\\frac{e^{-\\frac{k_T^{-2\\lambda}}{l \\lambda}}}{x}\\right)$ present in the general solution \\eqref{26} applying the initial distribution \\eqref{32}.\nLet us rewrite \\eqref{27} which is the rearranged form of the general solution \\eqref{26}\n\n\n\\begin{widetext}\n\\begin{equation}\n\t\\label{33}\n\t\\begin{split}\n\t\t\\text{G}\\left(\\frac{e^{-\\frac{k_T^{-2\\lambda}}{l \\lambda}}}{x}\\right)=&\\frac{1}{N_c^2-1}\\Bigg[x^{-m}N_c^2 \\bigg(\\frac{18\\alpha_s ^2 (-1)^{\\frac{1}{\\lambda }+1} \\lambda ^{1\/\\lambda } l^{1\/\\lambda } e^{-\\frac{m k_T^{-2\\lambda }}{\\lambda l}} m^{-\\frac{\\lambda l+l+n}{\\lambda l}} \\Gamma \\left(\\frac{l+n}{l \\lambda }+1,-\\frac{k_T^{-2\\lambda } m}{l \\lambda }\\right)}{\\pi R^2}\\\\\n\t\t&+\\frac{1}{f(x,k_T^2)} k_T^{-2\\frac{n}{l}} (-1)^{\\frac{n}{\\lambda l}} l^{-\\frac{n}{\\lambda l}} \\lambda ^{-\\frac{n}{\\lambda l}}\\bigg)+\\frac{1}{f(x,k_T^2)}x^{-m} k_T^{-2\\frac{n}{l}} (-1)^{\\frac{n}{\\lambda l}+1} l^{-\\frac{n}{\\lambda l}} \\lambda ^{-\\frac{n}{\\lambda l}}\\Bigg].\n\t\\end{split}\n\\end{equation}\n\\end{widetext}\nSetting initial parameters at $(x_0,k_T^2)$ we get initial distribution \\eqref{32} as\n\\begin{equation}\n\t\\label{34}\n\tf(x_0,k_T^2)=\\beta \\frac{x_0^{-\\lambda}\\sqrt{k_T^2}}{\\sqrt{\\ln\\frac{1}{x_0}}}\\exp \\left(-\\frac{\\ln^2(k_T^2\/k_s^2)}{2\\Omega\\ln (1\/x_0)}\\right).\n\\end{equation}\n\nWe denote the argument of G at $(x_0,k_T^2)$ as $\\tau=\\frac{e^{-\\frac{k_T^{-2\\lambda}}{l \\lambda}}}{x_0}$ which implies $k_T^2=( -l\\lambda \\ln (\\tau x_0))^{-1\/\\lambda }$. Now writing \\eqref{33} for $(x_0,k_T^2)$ we obtain\n\\begin{widetext}\n\\begin{equation}\n\t\\label{35}\n\t\\begin{split}\n\t\t\\text{G}(\\tau)=&x^{-m} k_T^{-\\frac{n}{l}} (-l\\lambda)^{\\frac{n}{\\lambda l}} \\Bigg(\\frac{x^{\\lambda } \\sqrt{\\ln \\frac{1}{x}} (\\lambda (-l) \\ln (\\tau x_0))^{1\/\\lambda } \\exp \\left(\\frac{\\ln ^2\\left((\\lambda (-l) \\ln (\\tau x_0))^{-1\/\\lambda }\\right)}{2 \\Omega \\ln \\frac{1}{x}}\\right)}{\\beta }\\\\\n\t\t&-\\frac{18\\alpha_s ^2 N_c^2 \\left(k_T^{-\\lambda }\\right)^{1\/\\lambda } \\left(e^{-\\frac{k_T^{-\\lambda }}{\\lambda l}}\\right)^m \\left(-\\frac{m k_T^{-\\lambda }}{\\lambda l}\\right)^{-\\frac{l+n}{\\lambda l}} \\Gamma \\left(\\frac{l+n}{l \\lambda }+1,-\\frac{k_T^{-\\lambda } m}{l \\lambda }\\right)}{\\pi m R^2 \\left(N_c^2-1\\right)}\\Bigg),\n\t\\end{split}\n\\end{equation}\n\\end{widetext}\nwhere, $k_T^2=( -l\\lambda \\ln (\\tau x_0))^{-1\/\\lambda }$. Note that \\eqref{35} is the functional form of G. We substitute $\\tau=\\frac{e^{-\\frac{k_T^{-2\\lambda}}{l \\lambda}}}{x}$ in \\eqref{35} which gives us the l.h.s. of \\eqref{33} and then solve \\eqref{33} for $f(x,k_T^2)$,\n\\begin{widetext}\n\\begin{equation}\n\t\\label{36}\n\t\\begin{split}\n\t\tf(x,k_T^2)=\\frac{\\beta m q^{n\/l} \\left(-\\frac{m k_T^{-2\\lambda }}{\\lambda l}\\right)^{\\frac{l+n}{\\lambda l}} \\left(-\\frac{m q^{-\\lambda }}{\\lambda l}\\right)^{\\frac{l+n}{\\lambda l}}}\n\t\t{\\left(-\\frac{m q^{-\\lambda }}{\\lambda l}\\right)^{\\frac{l+n}{\\lambda l}} \\left(\\tilde{\\Delta} q^{n\/l} \\tilde{A}[ k_T^2 ] +\\chi\\right)-\\tilde{\\Delta} \\beta k_T^{2n\/l} \\tilde{A}(q) \\left(-\\frac{m k_T^{-2\\lambda }}{\\lambda l}\\right)^{\\frac{l+n}{\\lambda l}}},\n\t\\end{split}\n\\end{equation}\n\n\n\nwhere,\n\\begin{equation*}\n\t\\begin{split}\n\t\t&\\chi=m x^{\\lambda } \\sqrt{\\ln \\frac{1}{x}} k_T^{2n\/l} e^{\\frac{\\ln ^2(q)}{2 \\Omega \\ln \\frac{1}{x}}} \\left(-\\frac{m k_T^{-2\\lambda }}{\\lambda l}\\right)^{\\frac{l+n}{\\lambda l}} q^{-1},\\text{ } \\text{ }\\tilde{A}[i]= i^{-1}\\beta \\left(e^{-\\frac{q^{-\\lambda }}{\\lambda l}}\\right)^m \\Gamma \\left[i\\right], \\text{ }\\text{ } \\text{ }\n\t\tq=\\left(-l\\lambda \\ln \\frac{x_0 e^{-\\frac{k_T^{-2\\lambda }}{\\lambda l}}}{x}\\right)^{-1\/\\lambda },\\\\\n\t\t&\\Gamma \\left[i\\right]=\\Gamma\\left(1+\\frac{l+n}{l\\lambda},\\frac{m (i)^{-\\lambda}}{l\\lambda}\\right), \\text{ }\\tilde{\\Delta}=\\frac{18 \\alpha _s^2}{\\pi R^2}\\frac{N_c^2}{N_c^2-1}.\n\t\\end{split}\n\\end{equation*}\n\\end{widetext}\n\\begin{figure}[t]\n\t\\centering\n\t\\includegraphics[width=.40\\textwidth,clip]{MD-BFKL3D1.eps}\n\t\\caption{\\label{5x} Three dimensional representation of (a) TMDlib HERA data fit: PB-NLO-HERAI+II+2018 (left) (b) our solution for KC improved MD-BFKL in $x$-$k_T^2$ phase space (right).}\n\\end{figure}\n\n\n\nFor simplicity let us denote r.h.s. of \\eqref{36} by $\\gamma$ i.e.\n\n\\begin{equation}\n\t\\label{37}\n\tf(x,k_T^2) = \\gamma(x,k_T^2).\n\\end{equation}\nAt $x=x_0$ and $k_T^2=k_0^2$\n\\begin{equation}\n\t\\label{38}\n\tf(x_0,k_0^2) = \\gamma(x_0,k_0^2).\n\\end{equation}\nNow dividing \\eqref{37} by \\eqref{38} we get\n\n\\begin{equation}\n\t\\label{39}\n\tf(x,k_T^2) = f(x_0,k_0^2)\\frac{\\gamma(x,k_T^2)}{\\gamma(x_0,k_0^2)}.\n\\end{equation}\nFrom \\eqref{32} we have the input distribution\n\\begin{equation}\n\t\\label{40}\n\tf(x_0,k_0^2)=\\beta \\frac{x_0^{-\\lambda}\\sqrt{k_0^2}}{\\sqrt{\\ln\\frac{1}{x_0}}}\\exp \\left(-\\frac{\\ln^2(k_0^2\/k_s^2)}{2\\Omega\\ln (1\/x_0)}\\right).\n\\end{equation}\n\\begin{figure*}[t]\n\t\\centering\n\t\\includegraphics[width=.29\\textwidth,clip]{lambda04.eps}\\hspace{1cm}\n\t\\includegraphics[width=.29\\textwidth,clip]{lambda06.eps}\n\t\\caption{\\label{6y} Density plots showing $\\lambda$ sensitivity of unintegrated gluon distribution $f(x,k_T^2)$, sketched for two canonical choice of $\\lambda$ viz. $\\lambda=0.4$ (left) and $\\lambda=0.6$ (right). }\n\\end{figure*}\nNow substituting $f(x,k_0^2)$ from \\eqref{40} into \\eqref{39} we obtain\n\\begin{equation}\n\t\\label{41}\n\t\\begin{split}\n\t\t&f(x,k_T^2)=\\\\&\\frac{(\\ln \\frac{1}{x_0})^{-\\frac{1}{2}}q^{n\/l} k_0^{-2\\frac{n}{l}+4\\frac{l+n}{ l}-2}\\beta x_0^{-\\lambda } e^{-\\frac{\\ln ^2(k_0^2)}{2 \\Omega \\ln \\frac{1}{x_0}}} \\left(k_T^2q\\right)^{-\\frac{l+n}{l}} \\tilde{\\phi}}\n\t\t{\\left(-\\frac{m q^{-\\lambda }}{\\lambda l}\\right)^{\\frac{l+n}{\\lambda l}} \\left(\\tilde{\\Delta} q^{n\/l} \\tilde{A}[k_T^2] +\\chi\\right)-\\tilde{\\Delta} k_T^{2n\/l} \\tilde{A}[q] \\left(-\\frac{m k_T^{-2\\lambda }}{\\lambda l}\\right)^{\\frac{l+n}{\\lambda l}}},\n\t\\end{split}\n\\end{equation}\nwhere\n\\begin{equation*}\n\t\\begin{split}\n\t\t\\delta=&m x_0^{\\lambda } \\sqrt{\\ln \\frac{1}{x_0}} k_0^{2n\/l-2} e^{\\frac{\\ln ^2(k_0^2)}{2 \\Omega \\ln \\frac{1}{x_0}}} \\left(-\\frac{m k_0^{-2\\lambda }}{\\lambda l}\\right)^{\\frac{l+n}{\\lambda l}}, \\\\\n\t\t\\tilde{\\phi}=&\\left(-\\frac{m k_0^{-2\\lambda }}{\\lambda l}\\right)^{\\frac{l+n}{\\lambda l}} \\left(\\tilde{\\Delta}k_0^{2n\/l} \\tilde{A}[k_0^2] +\\delta[x,k_T^2]\\right)\\\\\n\t\t&-\\tilde{\\Delta} k_0^{2n\/l}\\tilde{A}[k_0^2]\\left(-\\frac{m k_0^{-2\\lambda }}{\\lambda l}\\right)^{\\frac{l+n}{\\lambda l}}.\t\n\t\\end{split}\n\\end{equation*}\n\n\\begin{figure*}[tbp]\n\t\\centering\n\t\\includegraphics[width=.29\\textwidth,clip]{R5.eps}\\hspace{1cm}\n\t\\includegraphics[width=.29\\textwidth,clip]{R2.eps}\n\t\\caption{\\label{6xx} Density plots showing $R$ sensitivity of unintegrated gluon distribution $f(x,k_T^2)$, sketched for two form of shadowing: conventional $R=5 \\text{ GeV}^{-1}$ (left) and extreme $R=2 \\text{ GeV}^{-1}$ (right).} \n\\end{figure*}\n\nEquation \\eqref{41} serves as the parametric form of the solution for KC improved MD-BFKL equation. The input distribution is inclusive in the solution \\eqref{41} itself, therefore, we do not have to depend on the experimental data fits for input distribution unlike we did in the previous section of $x$ and $k_T^2$ evolution. This parametric form of the solution actually helps us to explore the three-dimensional insight of the gluon evolution in $x$-$k_T^2$ phase space. However, using \\eqref{41} one may also study $x$ and $k_T^2$ evolution separately by setting $x$ fixed for $k_T^2$ evolution or $k_T^2$ fixed for $x$ evolution.\n\n\n\\par In Fig.~\\ref{5x} we have shown our solution of KC improved MD-BFKL equation \\eqref{41} in three dimension. The kinematic region for our study is set to be $10^{-6}\\leq x \\leq10^{-2}$ and $1 \\text{ GeV}^2\\leq k_T^2 \\leq10^3\\text{ GeV}^2$. In the 3D surface, the suppression in the rise of the gluon distribution towards small $x$ due to shadowing correction is visible. However, a linear rise of the surface in the $k_T^2$ direction can be seen, attributed to the $1\/k_T^2$ factor in the nonlinear term, which offset the effect of shadowing at large $x$. \n\n\\par In Fig.~\\ref{6y} we have shown density plots of $f(x,k_T^2)$ in $x\\text{-}k_T^2$ domain to examine the sensitivity of $f(x,k_T^2)$ towards the parameter $\\lambda$. Density plot allows us to visualize and distinguish the kinematic regions with high\/low $f(x,k_T^2)$ in $x\\text{-}k_T^2$ plane which is more informative then any ordinary 3D plot. In Fig.~\\ref{6y} the plots are sketched for two canonical choices of $\\lambda$ viz. $\\lambda=0.4$ and $0.6$ corresponding to two $\\alpha_s$ values 0.15 and 0.23.\nOur solution seems to be very sensitive towards a small change in $\\lambda$. An apparent 50\\% change in $\\lambda$ (0.4 to 0.6) suggests approximately around one order of magnitude rise in gluon distribution $f(x,k_T^2)$ for an approximate limit of $x$ and $k_T^2$: $10^{-6}\\leq x\\leq 10^{-5}$ and $50 \\text{ GeV}^2 \\leq k_T^2\\leq 100 \\text{ GeV}^2$. It is also observed that the range of high $k_T^2$ and very small-$x$ is the high gluon distribution $f(x,k_T^2)$ region where gluons are mostly populated. In Fig.~\\ref{6xx} we have shown $R$ sensitivity of our solution for two choices of the shadowing parameter $R$ viz. $R= 5 \\text{ GeV}^{-1}$ and $2 \\text{ GeV}^{-1}$. A satisfactory shadowing effect is observed from the comparison of the two plots. The extreme form of shadowing $(R=2 \\text{ GeV}^{-1})$ is found to suppress atleast $10\\text{-}20 \\%$ magnitude of gluon density than the conventional shadowing $(R=5 \\text{ GeV}^{-1})$ in the high $k_T^2$ and small-$x$ region.\n\n\\section{Equation of Critical line and prediction of saturation scale: a differential geometric approach}\\label{critical}\n\\begin{figure*}[tbp]\n\t\\centering\n\t\\includegraphics[width=.4\\textwidth,clip]{Saturation_scale.eps}\\hspace{1.5cm}\n\t\\includegraphics[width=.4\\textwidth,clip]{critical.eps}\n\t\\caption{\\label{6x} (a) Contour plot obtained by solving \\eqref{35x} (left). Solid lines show contours of constant gradients along the curve. (b)Diagram showing critical boundary line separating high and low gluon density region (right). Blue dashed line is obtained from \\eqref{38x} while red dashed line is corresponding to GBW model \\cite{21}. }\n\\end{figure*}\nThe so-called saturation physics allows one to study the high parton density region in the small coupling regime. The transition from the linear region to saturation region is characterized by the saturation scale. The saturation momentum scale $Q_s$ is the threshold transverse momentum for which non-linearity becomes visible. The boundary in $(x, Q^2)$ or $(x,k_T^2)$ plane along which saturation sets in is characterized by the critical line. An important feature of our analytical solution to the KC improved MD-BFKL equation is the finding of the equation of the critical line which is supposed to mark the boundary between dilute and dense partonic system in $x\\text{-}k_T^2$ phase space. \n\\par Although the gluon saturation can be achieved only when $Q_s\\sim Q$ (or $k_T^2$), the observables already become sensitive to $Q_s$ during the approach to saturation regime. This property is known as geometrical scaling in DIS inclusive event which means that instead of depending on $k_T^2$ and $x$ separately, the gluon distribution depends on a single dimensionless variable $\\frac{k_T^2}{Q_s^2}$ i.e.\n\\begin{equation}\n\t\\label{31s}\n\t\\phi(x,k_T^2)\\equiv\\phi\\left(\\frac{k_T^2}{Q_s^2}\\right).\n\\end{equation}\nIn recent years, this geometrical scaling property of DIS observables is studied very extensively for various frameworks \\cite{42,56, 57, 58,84}. In \\cite{56} an analysis of the saturation scale has been performed in the platform of resummed NLLx BFKL where the saturation scale was calculated via the relation\n\\begin{equation}\n\t\\label{32s}\n\t-\\frac{d\\omega(\\gamma_c)}{d\\gamma_c}=\\frac{\\omega_s(\\gamma_c)}{1-\\gamma_c},\n\\end{equation}\nwhich has been repeatedly derived in several literature \\cite{59,60,61}. In \\cite{42} the saturation scale $Q_s$ was obtained from the numerical solution of a nonlinear equation by finding the maximum of the momentum distribution of the gluon. Another approach for determination of $Q_s$ can be found in \\cite{5} where a parameter $\\beta$ is defined as the relative difference between the solutions to the linear and nonlinear equation,\n\\begin{equation}\n\t\\label{33s}\n\t\\beta=\\frac{|f^\\text{lin}(x,Q_s^2(x,\\beta))-f^\\text{lin}(x,Q_s^2(x,\\beta))|}{f^\\text{lin}(x,Q_s^2(x,\\beta))}.\n\\end{equation} \nThe crucial parameter $\\beta$ actually depicts the percentage deviation that the non linear solution shows from the linear one and it lies in the order of $0.1 \\text{-} 0.5$ (or $10\\text{-}50 \\%$ deviation).\n\n\n\\par Our approach towards studying geometrical scaling and critical line is primarily based on the basic understanding from differential geometry, in particular gradient of a function which is considered as the direction of steepest ascent of that function. Each component of the gradient gives us the rate of change of the function with respect to some standard basis i.e. it gives us an idea about how fast our function grows or decays or saturates with respect to the change of the variables. One important advantage of choosing gradient is that it is a two dimensional object since it does not possess any component along $f(x,k_T^2)$ axis in $\\mathbb{R}^3$. This ensures that it does not have any direct dependence on the magnitude of gluon density $f(x,k_T^2)$, rather it depends on the rate at which $f(x,k_T^2)$ changes with respect to $x$ and $k_T^2$ change. This property of gradient actually helps us in distinguishing out the saturation region and linear region although the distribution function $f(x,k_T^2)$ has large variation in order of magnitude for different regions in $x\\text{-}k_T^2$ phase space.\n\\par Recall that towards small-$x$, gluon evolution is suppressed due to shadowing effect as sketch in Fig.~\\ref{5x} which motivates us to evaluate the gradient of $f(x,k_T^2)$ particularly along $x$ basis. For simplicity we consider an unit vector $\\vec{\\nu}$ along $\\tilde{\\eta} \\text{ }( = 1\/x)$ basis, we can project the gradient $\\nabla f(\\tilde{\\eta},k_T^2)$ along $\\vec{\\nu}$ via dot product $\\nabla f(\\tilde{\\eta},k_T^2)$.$\\vec{\\nu}$. This scalar quantity can also be interpreted as the directional derivative along the direction $\\vec{\\nu}$,\n\\begin{equation}\n\t\\label{34x}\n\tg\\equiv \\nabla_{\\vec{\\nu}} f(\\tilde{\\eta},k_T^2) = \\nabla f(\\tilde{\\eta},k_T^2).\\vec{\\nu}.\n\\end{equation}\nTaking the Euclidean norm yields\n\\begin{equation}\n\t\\label{35x}\n\tg = \\pm|\\nabla_{\\vec{\\nu}} f(\\tilde{\\eta},k_T^2)|,\n\\end{equation}\nwhere the negative (-) sign is for the descending function (or negative slope).\n\\par We obtained a family of contours (or level curves) $\\tilde{\\eta}(k_T^2)$ in $\\tilde{\\eta}\\text{-}k_T^2$ plane solving \\eqref{35x} as shown in Fig.~\\ref{6x}(a). Each contour depicts a constant gradient $g$ along the curves. The set of contours can also be identified as some set of possible saturation scales. As sketch in Fig.~\\ref{6x}(a) the $\\tilde{\\eta}\\text{-}k_T^2$ plane is divided into two regions: low gluon density region where the spacing between two consecutive contours is very small and high gluon density region where the spacing becomes very large compared to previous. This distinction in contour spacing in the two regions comes from the fact that the gradient changes very fast until the saturation is reached and then after reaching the saturation boundary gradient changes very slowly or almost freezes for further increase in $\\tilde{\\eta}(k_T^2)$ as can be seen in Fig.~\\ref{6x}(a). In high gluon density region, the contour curves tend to $\\tilde{\\eta}(k_T^2)\\rightarrow \\infty $ towards high $k_T^2$. For low gluon density region, shadowing effects are negligible and the contours become almost parallel straight lines. \n\n\n\\par Let us try to find the equation of critical line which divides the two regions in $\\tilde{\\eta}\\text{-}k_T^2$ space. The level curves of the function $g=\\pm| \\nabla_{\\vec{\\nu}} f(\\tilde{\\eta},k_T^2)| $ are two dimensional curves that can be obtained by setting $g=k$ where $k$ is a constant $(k\\in \\mathbb{R})$. Therefore, the equations of the level curves are given by \n\\begin{equation}\n\t\\label{36x}\n\t| \\nabla_{\\vec{\\nu}} f(\\tilde{\\eta},k_T^2)|=\\pm k.\n\\end{equation}\nNow for a known initial saturation momentum scale $Q_{s0}(1\/x_0)$ we can predict equation of the critical line using \\eqref{35x} and \\eqref{36x}. The equation of the critical line is the equation for the level curve of the function $g=\\pm| \\nabla_{\\vec{\\nu}} f(\\tilde{\\eta},k_T^2)| $ that passes through the point $(Q_{s0}(\\tilde{\\eta}_0),\\tilde{\\eta}_0)$. First we find the value of $k$ by plugging the point $(Q_{s0}(\\tilde{\\eta}_0),\\tilde{\\eta}_0)$ into \\eqref{36x}\n\\begin{equation}\n\t\\label{37x}\n\tk_{s0}=\\pm| \\nabla_{\\vec{\\nu}} f(Q_{s0}^2(\\tilde{\\eta}_0))|.\n\\end{equation}\nNow the level curve passing through $(Q_{s0}(\\tilde{\\eta}_0),\\tilde{\\eta}_0)$ is obtained by setting\n\\begin{equation}\n\t\\label{38x}\n\t| \\nabla_{\\vec{\\nu}} f(\\tilde{\\eta},Q_s^2)|=| \\nabla_{\\vec{\\nu}} f(Q_{s0}^2(\\tilde{\\eta}_0))|,\n\\end{equation}\nwhich is the equation of the critical boundary. The knowledge of an appropriate initial saturation scale $Q_{s0}(1\/x_0)$ allows one to separate out the linear and saturation region using \\eqref{38x}. In Fig.~\\ref{6x}(b) we have sketched a possible critical line obtained from \\eqref{38x} for the choice of initial saturation scale $Q_{s0}^2(\\eta_0)\\backsimeq 2.8 \\text{ GeV}^{2}$ at $\\eta_0=10^6$ (or $x_0=10^{-6}$) which is taken from the calculation from the original saturation model by Golec-Biernat and Wusthoff \\cite{21, 5}. A rough agreement between our prediction and that of GBW model is observed in Fig.~\\ref{6x}(b). However, $Q_s$ given by \\eqref{38x} is found to have weaker $x$ dependence than the one from GBW model $Q_s^{'2}$. The saturation scale has direct dependence on partons per unit transverse area. Smaller $x$ suggests larger parton density giving rise to a larger saturation momentum scale, $Q_s^2$. In other words the saturation scale $Q_s$ depends on $x$ in such a way that with decreasing $x$ one has to probe smaller distances or higher $Q^2$ in order to resolve the dense partonic structure of the proton which is clear from our analysis.\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\\section{KC improved MD-BFKL prediction of HERA DIS data}\\label{hera}\n\\subsection{DIS structure functions and reduced cross section}\\label{hera1}\nIn this section we present a quantitative prediction of proton structure functions $F_2(x,Q^2)$ and longitudinal structure function $F_L(x,Q^2)$ as an outcome of our solution to the KC improved MD-BFKL equation. At small-$x$, the sea quark distribution is driven by gluons via. $g\\rightarrow q\\bar{q}$ process. This component from sea quark distribution can be calculated in perturbative QCD. The relevant diagram for this QCD process is shown in Fig.~\\ref {d1} and the contribution to the transverse and longitudinal components of the structure functions can be written using the $k_T$-factorization theorem \\cite{25,26}\n\n\\begin{equation}\n\t\\label{42}\n\tF_\\text{i}(x,Q^2)=\\int_x^1\\frac{dx^{'}}{x^{'}}\\int\\frac{dk_T^2}{k_T^4}f\\left(\\frac{x}{x^{'}},k_T^2\\right)F_\\text{i}^{(0)}(x^{'},k_T^2,Q^2),\n\\end{equation}\nwhere i= T, L and $\\frac{x}{x^{'}}$ is the fractional momentum carried by gluon which splits into $q\\bar{q}$ pair. $F_\\text{i}^{(0)}$ includes both the quark box and cross box contribution which comes from virtual gluon-virtual photon subprocess leading to $q\\bar{q}$ production (Fig.~\\ref{d1} ). The gluon distribution $f(\\frac{x}{x^{'}},k_T^2)$ in \\eqref{42} represents the sum of the gluon ladder diagrams in the lower part of the box as shown in Fig.~\\ref {d1} is given by BFKL equation \\cite{2}. Here we will study the effect of gluon shadowing using the solution $f(x,k_T^2)$ of KC improved MD-BFKL in \\eqref{42}.\n\n\n\\par The explicit expression for quark box contribution $F_\\text{i}^{0}$ can be obtained by writing four momentum in terms of the convenient light like momenta $p$ and $q^{'}\\equiv q +xp$, where $x=\\frac{Q^2}{2p.q}$ and $Q^2=-q^2$ like as usual (see Fig.~\\ref{d1} ). Now we can decompose quark and gluon momentum in terms of sudakov variables\n\\begin{equation}\n\t\\begin{split}\n\t\t&\\kappa=\\alpha p-\\beta q^{'}+\\kappa_T,\\\\\n\t\t&k= a p+bq^{'}+k_T.\n\t\\end{split}\n\\end{equation}\nThe integration should be performed over the box diagram subject to quark mass-shell constraints \\cite{8}\n\n\\begin{align}\n\t\\begin{split}\n\t\t&(\\alpha -x)2p.q(1-\\beta)-\\kappa_T^2=m_q^2,\\\\\n\t\t&(a-\\alpha)2p.q\\beta-(\\kappa_T-k_T)^2=m_q^2,\n\t\\end{split}\n\\end{align}\nwhich leads to the box contribution \\cite{29}\n\\begin{align}\n\t\\label{43}\n\t\\begin{split}\n\t\t&\\tilde{F}_T^{(0)}(k_T^2,Q^2)=2 \\sum_q e_q^2\\frac{Q^2}{4\\pi^2}\\int_{0}^{1}d\\beta\\int d^2\\kappa_T\\alpha_s(\\kappa_T)\\\\&\\times\\big\\{[\\beta^2+(1-\\beta)^2]\\bigg[ \\frac{\\kappa_T^2}{L_1^2}-\\frac{\\kappa_T.(\\kappa_T-k_T)}{L_1L_2}\\bigg]+\\frac{m_q}{L_1^2}-\\frac{m_q^2}{L_1L_2} \\big\\},\n\t\t\\end{split}\n\t\\end{align}\n\t\\begin{align}\n\t\\label{44}\n\t\\begin{split}\n\t\t\\tilde{F}_L^{(0)}(k_T^2,Q^2)=&2 \\sum_q e_q^2\\frac{Q^4}{\\pi^2}\\int_{0}^{1}d\\beta\\int d^2\\kappa_T\\alpha_s(\\kappa_T)\\\\&\\times\\beta^2(1-\\beta)^2\\bigg( \\frac{1}{L_1^2}-\\frac{1}{L_1L_2}\\bigg),\n\t\t\\end{split}\n\\end{align}\nwhere the denominators $L_\\text{i}$ are\n\\begin{equation*}\n\t\\begin{split}\n\t\t&L_1=\\kappa_T^2+\\beta(1-\\beta)Q^2+m_q^2,\\\\\n\t\t&L_2=(\\kappa_T-k_T)^2+\\beta(1-\\beta)Q^2+m_q^2.\n\t\\end{split}\n\\end{equation*}\n\\begin{figure*}[tbp]\n\t\\centering\n\t\\includegraphics[width=.23\\textwidth,clip]{Structure_box1.eps}\n\t\\hspace{2cm}\n\t\\includegraphics[width=.23\\textwidth,clip]{Structure_Cross1.eps}\n\t\n\n\t\\caption{\\label{d1} Diagrammatic representation of the factorization formula \\eqref{42} where gluon couples to virtual photon through the (a) quark box (left) and (b) crossed box (right) diagrams. }\n\\end{figure*}\n\n\nNote that $\\tilde{F}_\\text{i}^{(0)}(k_T^2,Q^2)\\equiv \\int_x^1\\frac{dx^{'}}{x^{'}}F_\\text{i}^{(0)}(x^{'},k_T^2,Q^2)$ i.e. the $x^{'}$ integration of \\eqref{42} is implicit in the $d^2k_T^{'}$ and $d\\beta$ integration in \\eqref{43} and \\eqref{44}. Now plugging \\eqref{43} and \\eqref{44} in the $k_T$-factorization formula \\eqref{42} we get\n\\begin{align}\n\t\\label{45}\n\t\\begin{split}\n\t\tF_T(x,Q^2)=&2 \\sum_q e_q^2\\frac{Q^2}{4\\pi^2}\\int_{k_0^2}^\\infty\\frac{dk_T^2}{k_T^4} \\int_{0}^{1}d\\beta\\int d^2\\kappa_T\\alpha_s(\\kappa_T)\\\\&\\times\\big\\{[\\beta^2+(1-\\beta)^2]\\bigg[ \\frac{\\kappa_T^2}{L_1^2}-\\frac{\\kappa_T.(\\kappa_T-k_T)}{L_1L_2}\\bigg]\\\\\n\t\t&+\\frac{m_q}{L_1^2}-\\frac{m_q^2}{L_1L_2} \\big\\}f(\\frac{x}{x^{'}},k_T^2),\n\t\t\\end{split}\n\t\\end{align}\n\\begin{align}\n\\begin{split}\t\n\t\\label{46}\n\t\tF_L(k_T^2,Q^2)=&2 \\sum_q e_q^2\\frac{Q^4}{\\pi^2}\\int_{k_0^2}^\\infty\\frac{dk_T^2}{k_T^4}\\int_{0}^{1}d\\beta\\int d^2\\kappa_T\\alpha_s(\\kappa_T)\\\\&\\times\\beta^2(1-\\beta)^2\\bigg( \\frac{1}{L_1^2}-\\frac{1}{L_1L_2}\\bigg)f(\\frac{x}{x^{'}},k_T^2).\n\t\\end{split}\n\\end{align}\n\\begin{figure*}[tbp]\n\t\\label{}\n\t\\centering\n\t\\includegraphics[trim=21 0 21.5 0,width=.45\\textwidth,clip]{F2_5.eps}\\hspace{5mm}\n\t\\includegraphics[trim=21 0 21.5 0,width=.45\\textwidth,clip]{F2_15.eps}\\\\\n\t\\includegraphics[trim=21 0 21.5 0,width=.45\\textwidth,clip]{F2_35.eps}\\hspace{5mm}\n\t\\includegraphics[trim=21 0 21.5 0,width=.45\\textwidth,clip]{F2_45.eps}\n\t\n\t\n\t\n\n\t\\caption{\\label{7x} Prediction for proton structure function F$_2$ obtained for two choices of shadowing: conventional ($R= 5 \\text{ GeV}^{-1}$) and \"hotspot\" ($R= 2 \\text{ GeV}^{-1}$). Data are taken from HERA (H1 \\cite{32} and ZEUS \\cite{33}) as well as fixed target experiment (NMC \\cite{31} and BCDMS \\cite{34}). Data-sets from global parametrization groups NNPDF3.1sx \\cite{36} and PDF4LHC 15 \\cite{35} is also included. The background contribution is given by \\eqref{49} with $F_2^{BG}(x_0=0.1)\\approx$ 0.316, 0.384, 0.391 and 0.406 corresponding to four $Q^2$ values viz. $5 \\text{ GeV}^2, 15 \\text{ GeV}^2, 35 \\text{ GeV}^2$ and $45 \\text{ GeV}^2$. Separate plots of $F_2^{\\text{npert.}}$ vs. $F_2^{\\text{npert.+pert.}}$ is also sketched.\n\t}\n\\end{figure*}\nEquations \\eqref{45} and \\eqref{46} serve as the basic tool for calculating structure functions at small-$x$ provided that the gluon distribution $f(x,k_T^2)$ is known. An analytical approach towards the calculation of $F_2$ and $F_L$ can be found in literature \\cite{27,28} for fixed coupling case using linear BFKL equation. In the literature \\cite{9}, structure functions are calculated taking gluon distribution $f(x,k_T^2)$ from the numerical solution of unitarized BFKL equation for running coupling consideration. It has been seen that the analytical approach \\cite{28} considerably overestimates the actual (numerical) solution \\cite{9}. This is because of the fact that analytical approach neglects terms down only by powers of $\\ln (1\/z)^{-1}$ as well as it does not accommodate running of strong coupling. However, our approach is a semi-analytical one, in the sense that we take gluon distribution from our analytical solution while further integrations are performed numerically. This indeed allows us to check the feasibility of our analytical solution in describing DIS data. \n\\par In \\eqref{42} $m_q$ denotes the quark mass and it is taken to be $m_q=1.28 \\text{ GeV}$ for charm quark while massless ($m_q=0$) for light quarks (u, d and s). Our phenomenology is limited for light quark therefore putting $m=0$ in \\eqref{43} and replacing the quark transverse momentum by $\\kappa_T=\\kappa_T^{'}+(1-\\lambda)k_T$ we obtain \\cite{28}\n\\begin{equation}\n\t\\label{47}\n\t\\begin{split}\n\t\t&\\tilde{F}_T^{(0)}(k_T^2,Q^2)=\\\\& \\sum_q e_q^2\\frac{\\alpha_s}{\\pi}Q^2k_T^2\\int_{0}^{1}d\\beta\\int_{0}^{1}d\\lambda\\int_0^{\\infty} d\\kappa_T^{'2}[\\beta^2+(1-\\beta)^2]\\\\&\\times\\lambda\\frac{(2\\lambda-1)\\kappa_T^{'2}+(1-\\lambda)[\\lambda(1-\\lambda)k_T^2+\\beta(1-\\beta)Q^2]}{[\\kappa_T^{'2}+\\lambda(1-\\lambda)k_T^2+\\beta(1-\\beta)Q^2]^3}.\n\t\\end{split}\n\\end{equation}\nAfter integrating over $\\kappa_T^{'2}$ one can arrive at\n\\begin{equation}\n\t\\label{48}\n\t\\begin{split}\n\t\t\\tilde{F}_T^{(0)}(k_T^2,Q^2)=& \\sum_q e_q^2\\frac{\\alpha_s}{\\pi}Q^2k_T^2\\int_{0}^{1}d\\beta\\int_{0}^{1}d\\lambda\\\\&\\times\\frac{[\\lambda^2+(1-\\lambda)^2][\\beta^2+(1-\\beta)^2]}{\\lambda(1-\\lambda)k_T^2+\\beta(1-\\beta)Q^2}.\n\t\\end{split}\n\\end{equation}\nEquation \\eqref{47} and \\eqref{48} are written in terms of Feynman integral which actually eliminates the azimuthal dependence and reduces the two fold integral $d^2\\kappa_T$ of \\eqref{43} to single integral $\\pi d\\kappa_T^{'2}$ . From \\eqref{48} it is clear that $\\tilde{F}_T^{(0)}(k_T^2,Q^2)$ or $F(x^{'},k_T^2,Q^2)$ possess the dimension of $k_T^2$. Therefore, $F(x^{'},$$k_T^2,$ $Q^2)$ or more conveniently $F(x^{'},k_T^2,Q^2)\/k_T^2$ may be considered as the structure function of an off mass shell gluon of approximate virtuality $k_T^2$. In \\cite{27} differential structure functions have been studied for fixed coupling and it is found that the ratio between longitudinal $F_L$ and transverse structure function $F_T$ is 2:9 for fixed coupling approximation. We have considered this ratio directly in our calculation of longitudinal structure function. Finally, we have taken an assumption $f(x\/x^{'}.k_T^2)\\rightarrow f(x,k_T^2)$ i.e. we ignore the $x^{'}$ dependence of $f(x\/x^{'}.k_T^2)$ which is reasonable in LLx accuracy since \n\\begin{equation*}\n\t(\\ln\\frac{x}{x^{'}})^n=(\\ln x)^n[1+O(1\/\\ln x)].\n\\end{equation*}\nThe advantage of taking this assumption is that we do not have to impose the possible constraint \\cite{8} coming from $x\/x^{'}<1$ on the region of integration. In principle, the factorization formula \\eqref{42} require to be run down to $k_T^2=0$. The integral itself is infrared finite as both the functions $f(\\frac{x}{x^{'}},k_T^2)$ and $F_i^{(0)}(x^{'},k_T^2,Q^2)$ vanish at $k_T^2=0$. However, BFKL dynamics is based on perturbative QCD which is not expected to hold the nonperturbative small $k_T^2$ physics. On the other hand, for small $k_T^2$ the gluon distribution vanishes linearly with the decrease in $k_T^2$ on account of gauge invariance \\cite{9} making the contribution small. Therefore, we have neglected this small contribution from small $k_T^2$ region in our calculations of unintegrated gluon distribution $f(x,k_T^2)$.\n\n\\begin{figure*}[tbp]\n\t\\label{}\n\t\\centering\n\t\\includegraphics[trim=21 0 21.5 0,width=.45\\textwidth,clip]{FL_5.eps}\\hspace{5mm}\n\t\\includegraphics[trim=21 0 21.5 0,width=.45\\textwidth,clip]{FL_15.eps}\\\\\n\t\\includegraphics[trim=21 0 21.5 0,width=.45\\textwidth,clip]{FL_35.eps}\\hspace{5mm}\n\t\\includegraphics[trim=21 0 21.5 0,width=.45\\textwidth,clip]{FL_45.eps}\n\t\n\t\n\t\n\n\t\\caption{\\label{77x} Prediction for proton structure function $F_L$ obtained for two choices of shadowing: conventional ($R= 5 \\text{ GeV}^{-1}$) and \"hotspot\" ($R= 2 \\text{ GeV}^{-1}$). Data are taken from HERA H1 \\cite{49} and ZEUS \\cite{33}. Data-sets from global parametrization group PDF4LHC 15 \\cite{35} is also included. The background contribution is given by \\eqref{49} with $F_L^{BG}(x_0=0.1)\\approx$ 0.057, 0.063, 0.071 and 0.073 corresponding to four $Q^2$ values viz. 5 $\\text{ GeV}^2$, 15 $\\text{ GeV}^2$, 35 $\\text{ GeV}^2$ and 45 $\\text{ GeV}^2$. Separate plots of $F_2^{\\text{npert.}}$ vs. $F_2^{\\text{npert.+pert.}}$ is also sketched. }\n\\end{figure*}\n\\begin{figure}[tbp]\n\t\\centering\n\t\\includegraphics[width=.48\\textwidth,clip]{RQ1.eps}\n\t\\caption{\\label{Rx} pQCD prediction from KC improved MD-BFKL for the ratio $R(Q^2)$ in the kinematic range $2 \\text{ GeV}^2\\leq Q^2\\leq 100 \\text{ GeV}^2$. The data \\cite{49} are from H1 ($\\sqrt{s}= 252\\text{ GeV}$) and ZEUS ($\\sqrt{s}= 225\\text{ GeV}$) experiment. The theoretical calculation are for $\\sqrt{s}= 225\\text{ GeV}$ analogous to c.m.s. of ZEUS .\n\t}\t\n\\end{figure}\n\n\\par Before proceeding to the realistic estimation of the structure functions, we add on the \"background\" some non-BFKL contribution, $F_\\text{i}^{\\text{BG}}$ to $F_\\text{i}$ since the above prediction of structure function is not enough for describing DIS data \\cite{9}. This is because \\eqref{45} and \\eqref{46} represent only the LLx gluon contribution and they are not the only contributions to the DIS structure functions. Although gluonic contribution is dominant at small-$x$, towards higher $x$ their effect becomes weak and we cannot neglect other non-BFKL contribution. For instance, we assume that $F_\\text{i}^{\\text{BG}}$ evolves like $x^{-0.08}$ motivated from soft pomeron intercept $\\alpha_{P}(0)=1.08$ \\cite{30}. To be precise we use \n\\begin{equation}\n\t\\label{49}\n\tF_\\text{i}^{\\text{BG}}(x,Q^2)=F_\\text{i}(x_0,Q^2)\\left(\\frac{x}{x_0}\\right)^{-0.08}.\n\\end{equation}\nAn intelligent way of calculating this non-BFKL contribution is to choose $x_0$ at some high $x$ and then take $F_\\text{i}^{\\text{BG}}(x_0,Q^2)$ from data \\cite{31} which is also listed in the figure caption. The recent HERA DIS data taken for comparison with our result can be found in \\cite{32,49} from H1 collaboration and in \\cite{33} from ZEUS collaboration. In the text \\cite{32} from H1 collaboration inclusive neutral current $e^{\\pm}p$ scattering cross section data collected during the years (2003-2007) is presented. The beam energies $E_p$ of corresponding H1 experiment run are $920$, $575$ and $460 \\text{ GeV}^2$. Corresponding $F_L$ data of H1 is taken from \\cite{49} where measurement are performed at c.m.s. energies $\\sqrt{s}=$ 225 and 252 $\\text{GeV}$. On the other hand, in \\cite{33} from ZEUS collaboration reduced cross sections for $ep$ scattering for different c.m.s. energies viz. $318$, $251$ and $225\\text{ GeV}$ is presented. The fixed target data from NMC \\cite{31} and BCDMS \\cite{34} collaboration which exist for $x>10^{-2}$ is also shown in the Figure \\ref{7x}. Finally, we have included data from global parameterization groups viz. NNPDF3.1sx \\cite{36} and PDF4LHC15 \\cite{35} for comparison. \n\n\n\\begin{figure*}[tbp]\n\t\\centering\n\t\\includegraphics[width=.32\\textwidth,clip]{cros2.eps}\n\t\\includegraphics[width=.32\\textwidth,clip]{cros5.eps}\n\t\\includegraphics[width=.32\\textwidth,clip]{cros85.eps}\\\\\n\t\\includegraphics[width=.32\\textwidth,clip]{cros12.eps}\t\n\t\\includegraphics[width=.32\\textwidth,clip]{cros25.eps}\n\t\\includegraphics[width=.32\\textwidth,clip]{cros45.eps}\n\n\t\\caption{\\label{8x} Theoretical prediction from KC improved MD-BFKL for reduced cross section $\\sigma_r (x,Q^2)$. The data are from H1 \\cite{40,83} ($\\sqrt{s}\\approx$ 318, 300 and 252 $\\text{GeV}^2$) and ZEUS \\cite{33} ($\\sqrt{s}\\approx$ 318 $\\text{GeV}^2$). Theoretical calculations are for $\\sqrt{s}= 318 \\text{ GeV}^2$.}\n\\end{figure*}\n\\par The $x$ dependence of the structure functions $F_L$ and $F_2$ is shown in Fig.~\\ref{7x} and Fig.~\\ref{77x} for the four $Q^2$ values $5 \\text{ GeV}^2$, $15 \\text{ GeV}^2$, $35 \\text{ GeV}^2$ and $45 \\text{ GeV}^2$. The nonperturbative contribution, $F_{2\/L}^{\\text{npert.}}$ (or $F_{2\/L}^{\\text{BG}}$) given by \\eqref{49} contrasted with total (nonpert. +pert.) contribution, $F_{2\/L}^{\\text{npert.+pert.}}$. QCD prediction shows a satisfactory agreement with DIS data for both structure functions $F_2$ and $F_L$. It is clear that perturbative contribution is insignificant towards large $x$ ($\\geq 10^{-2}$), while for small $x$ around $x\\leq 10^{-3}$ this contribution becomes visibly important. Interestingly both data and theory for $F_2$ structure function seem to preserve the $x^{-\\lambda}$ singular behavior rather than showing taming due to net shadowing effect in the asymptotic limit $x\\rightarrow0$. It is difficult to observe the existence of any sizable shadowing effect for $x>10^{-3}$. However, for $x<10^{-3}$, recalling that the shadowing term is proportional to $1\/R^2$ it is seen that for $R=2\\text{ GeV}^{-1}$ (gluons in hotspots) the rise of structure function becomes slower than that of $R=5\\text{ GeV}^{-1}$ ($R=R_H$, hadron radius) as expected. But even the extreme form of the gluon shadowing ($R=2\\text{ GeV}^{-1}$) suppresses $F_2$ only by about $10\\%$ or less at $10^{-3}$. It just depicts that shadowing has the negligible impact on structure functions even towards smaller $x$. This is because of the fact that in this low $x$ regime gluons are expected to drive the sea quark distribution via $g\\rightarrow q\\bar{q}$. Therefore, a similar sea quark contribution to the structure function $F_2$ in addition to the gluon contribution can be expected in low $x$ regime. On the other hand from Fig.~\\ref{77x} it is clear that for $x>10^{-3}$, the size of the longitudinal structure function is negligibly small while for very small-$x$ regime ($x<10^{-3}$) , $F_L$ grows eventually. This is in accord with our expectation since measurement of $F_L$ directly probes the gluonic content of the proton which is dominant in small-$x$ regime. \n\n\n\\par The proton structure functions $F_2$ and $F_L$ are of complementary in nature. These are related to the $\\gamma^{*}p$ cross sections of longitudinally and transversely polarized photons $\\sigma_L$ and $\\sigma_T$ as $F_L\\propto\\sigma_L$ and $F_2\\propto (\\sigma_L+\\sigma_T)$. Since $\\sigma_L$ and $\\sigma_T$ are positive, that imposes a restriction on $F_L$ and $F_2$ i.e. $0\\leq F_L\\leq F_2$. To circumvent the need of a proper relationship between the structure functions and polarized cross sections to study $\\gamma^*p$ cross section one can define a ratio between the structure functions or the equivalent cross section ratio\n\\begin{equation}\n\t\\label{R}\n\tR=\\frac{\\sigma_L}{\\sigma_T}=\\frac{F_L}{F_2-F_L}.\n\\end{equation}\nThe advantage behind this formulation is that this is independent of any normalization factors.\n\n\\par In Fig \\ref{Rx} a phenomenological comparison between data and theory is shown to illustrate $Q^2$ dependence of the ratio $R(x,Q^2)$. The HERA data taken for comparison can be found in \\cite{49} where $ep$ cross section data measured for two center of mass energies $\\sqrt{s}=225$ and $252\\text{ GeV}$ is enlisted. From Fig.~\\ref{Rx} it is clear that the HERA data is in general agreement with the QCD expectation. Available H1 data for very low $Q^2$ $(Q^2\\leq 5\\text{ GeV})$ is also sketched in Fig.~\\ref{Rx}. We have excluded the very low $Q^2$ region in our calculation of $R$ since we are bounded to stick in the perturbative region. However, a more realistic study in this transition region is performed in Sect.~\\ref{hera2}. \n\n\\par Now we try to predict reduced cross section $\\sigma_r$ for $ep$ scattering process from the knowledge of structure functions that we have discussed above. The inclusive deep-inelastic differential cross section for $ep$ scattering can be represented in terms of three structure functions $F_2$, $F_L$ and $xF_3$. The structure functions have direct dependence with DIS kinematic variables, $x$ and $Q^2$ only, while the cross section has additional dependence with inelasticity $y=Q^2\/s x$. The inclusive cross section for neutral current $ep$ scattering is given by\n\\begin{equation}\n\t\\label{50}\n\t\\begin{split}\n\t\t\\frac{d^2 \\sigma^{e^{\\pm}p}}{dx dQ^2}=\\frac{2\\pi\\alpha^2}{x Q^4}Y_{+}\\left(F_2-\\frac{y^2}{Y_+}F_L\\mp\\frac{Y_{-}}{Y_{+}}xF_3\\right),\n\t\\end{split}\n\\end{equation}\nwhere $Y_{\\pm}=1\\pm(1-y)^2$. The cross section for exchanged virtual boson $(\\text{Z} \\text{ or } \\gamma^{*})$ is related to $F_2$ and $F_3$ in which contribution from both longitudinal and transverse boson polarization state exists. On the other hand, only the longitudinal polarized virtual boson exchange processes contribute to $F_L$ which has a significant impact on higher order QCD though it vanishes at lowest order. At small momentum transfer $Q^2$ $(\\text{i.e. } Q^2\\ll M_Z^2\\approx800\\text{ GeV}^2)$, interaction of massless photon is dominant over the exchange of heavy Z boson. Thus, the contribution from Z boson exchange and the influence of the term $xF_3$ is negligible at low and moderate $Q^2$. Therefore, in this range of $Q^2$, in one photon exchange approximation, the differential cross section formula \\eqref{50} can be written as \n\n\\begin{equation}\n\t\\label{51}\n\t\\sigma_r\\equiv\\frac{d^2 \\sigma^{e^{\\pm}p}}{dx dQ^2}\\frac{x Q^4}{2\\pi\\alpha^2}\\frac{1}{Y_{+}}=\\left(F_2-\\frac{y^2}{Y_+}F_L\\right),\n\\end{equation}\nwhere $\\sigma_r$ is the reduced cross section. Note that \\eqref{51} is symmetric under charge exchange i.e. identical for both $e^{+}p$ and $e^{-}p$ processes unlike \\eqref{50}. Additionally, it is independent of incoming electron helicity state. Thus, at low and moderate $Q^2$ $(\\leq m_Z^2)$ which is our region of study, the knowledge of $F_2$ and $F_L$ is enough to predict the reduced cross section. We can also express reduced cross section in terms of the ratio $R(x,Q^2)$ defined in \\eqref{R} replacing $F_L$ by $F_L=\\frac{R}{1+R}F_2$ which yields\n\\begin{equation}\n\t\\label{51x}\n\t\\sigma_r=F_2(x,Q^2)\\left[1-\\frac{y^2}{Y_+}\\frac{R}{1+R}\\right].\n\\end{equation}\n\n\\par The $x$ dependence of $e^{\\pm}p$ reduced cross section $\\sigma_r$ calculated for center of mass energy $\\sqrt{s}=318\\text{ GeV}$ is shown in Fig.~\\ref{8x}. Our theoretical expectation is compared with HERA H1 measurement \\cite{40,83} and ZEUS ($\\sqrt{s}=318\\text{ GeV}$) \\cite{33}. The available H1 data for low $Q^2$ ($\\leq12\\text{ GeV}^2$) measured at $\\sqrt{s}=$ 318 $\\text{GeV}$ (SVX) and 300 $\\text{GeV}$ (NVX-BST) is taken from \\cite{40}, while the same for high $Q^2$ ($>12\\text{ GeV}^2$) is taken from \\cite{83}.\nThe cross section measurement of SVX is found to be slightly higher than that of NVX-BST as expected because of the increase in center of mass energy. Both theory and data agree well for our phenomenology range $Q^2\\leq 100\\text{ GeV}^2$. The distinction in $\\sigma_r$ due to the two forms of shadowing $R=5\\text{ GeV}^{-1}$ and $R=2\\text{ GeV}^{-1}$ is more prominent for smaller $Q^2$ values. For each $Q^2$, starting at some high $x$ the reduced cross section first increases as $x\\rightarrow0$ and then an abrupt fall in cross section can be observed in both theory and data at very small-$x$ region ($x<10^{-4}$) . For all $Q^2$, this region of $x$ corresponds to the highest inelasticity $y\\approx0.65$ ($y=Q^2\/sx$) and thus characteristic turn over of cross section at $y\\approx0.65$ can be attributed to the influence of $F_L$. In simple words, towards very small-$x$ (or high $y$) the monotonic rise of $F_2$ is suppressed by the contribution from longitudinal structure function $F_L$ thereby causing an overall fall in the cross section. For low inelasticity $y<0.65$, the contribution from the longitudinal structure function is small on the other hand, structure function $F_2$ exhibits a steady increase as $x\\rightarrow 0$. Therefore, in the region where $x$ is not so small, the growth of the cross section is found to be power like as expected. \n\n\n\\subsection{Virtual photon-proton cross section prediction in transition region}\\label{hera2}\n\\par Traditionally the photon-proton interaction is classified into two separate processes depending upon the photon virtuality $Q^2$ viz. photoproduction (at low $Q^2$) and deep inelastic scattering (at high $Q^2$). DIS is considered as a basic tool for exploring pQCD where the point like virtual photon directly probes the partonic contents of the proton. On the other hand, photoproduction is completely nonperturbative phenomena defined in the limit of vanishing $Q^2$ where real (or quasi-real) photons interact with the proton more likely a hadron-hadron collision. The experiment at HERA collider provides a unique opportunity to study both the processes photoproduction and DIS on their respective kinematic domains. The $Q^2$ dependence of the proton structure functions are well described by pQCD over a wide range of $x$ and $Q^2$ \\cite{37,38} in accordance with HERA data. However, for $Q^2\\lesssim 2\\text{ GeV}^2$ (photo production region) the pQCD breakdowns since the higher order contributions to the perturbative expansion becomes very large. In this region, data can be only described by non-perturbative phenomenological models \\cite{39}. Our present study is especially focused on the transition region $(2\\text{ GeV}^2\\leq Q^2\\leq10\\text{ GeV}^2)$ from photoproduction to deep inelastic scattering. To measure the photon-proton cross section in the transition region two dedicated runs were performed in the years 1999 (Nominal vertex \"NVX'99\") and 2000 (Shifted Vertex \"SVX'00\") by H1 experiment at HERA. The published data can be found in \\cite{40}.\n\n\n\n\\par Recall the neutral current $ep$ double differential cross section formula \\eqref{51} defined in the region $Q^2\\ll M_Z^2$. In this region massive boson $(M_Z)$ exchange is neglected and only one photon exchange is considered, thereby the role of incoming electron reduces to be a source of virtual photon interacting proton. Thus, we can recast the formula \\eqref{51} in terms of photon-proton reaction. In fact the structure function $F_2$ and $F_L$ in \\eqref{51} related to the longitudinally and transversely polarized photon-proton scattering cross sections $\\sigma_L$ and $\\sigma_T$ by the relations\n\\begin{align}\n\t\\label{52}\n\t&F_L = \\frac{Q^2}{4\\pi^2\\alpha}(1-x)\\sigma_L,\\\\\n\t\\label{53}\n\t&F_2 = \\frac{Q^2}{4\\pi\\alpha}(1-x)(\\sigma_L+\\sigma_T),\n\\end{align}\nwhich hold good at low $x$. Considering \\eqref{52} and \\eqref{53} the reduce cross section in \\eqref{51} can be written as\n\\begin{equation}\n\t\\label{54}\n\t\\sigma_r = \\frac{Q^2(1-x)}{4\\pi^2\\alpha}\\sigma_{\\gamma^*p}^{\\text{eff}},\n\\end{equation}\nwhere \\begin{align}\n\t\\label{55}\n\t\\sigma_{\\gamma^*p}^{\\text{eff}}=\\sigma_T+(1-\\frac{y^2}{Y_+})\\sigma_L\n\\end{align}\nis the effective virtual photon-proton cross section. Note that the expression for $\\sigma_{\\gamma^*p}^{\\text{eff}}$ is similar to the total cross section $\\sigma_{\\gamma^*p}^{\\text{tot}}$ which is linear combination of $\\sigma_L$ and $\\sigma_T$ i.e. \n$\\sigma_{\\gamma^*p}^{\\text{tot}}=\\sigma_L+\\sigma_T$. In fact the total cross section $\\sigma_{\\gamma^{*}p}^{\\text{tot}}$ and $\\sigma_{\\gamma^*p}^{\\text{eff}}$ can be regarded as the same quantity at low inelasticity $y$ i.e. $\\sigma_{\\gamma^*p}^{\\text{eff}} \\xrightarrow{y\\rightarrow0}\\sigma_{\\gamma^*p}^{\\text{tot}}$ which differ only in the region of high y.\n\n\\begin{figure}[tbp]\n\t\\centering\n\t\\includegraphics[width=.4\\textwidth,clip]{W_cross.eps}\n\t\\caption{\\label{7xx} QCD prediction from KC improved MD-BFKL for virtual photon-proton cross section $\\sigma_{\\gamma^{*}p}^\\text{eff}$ as a function of $Q^2$ at different values of $W$. The cross section for different $W$ values are multiplied by factor multiple of 2 indicated in the figure. The included data are from H1 \\cite{40} ($\\sqrt{s}= 318\\text{ GeV}$) and ZEUS \\cite{81} ($\\sqrt{s}= 300\\text{ GeV}$). }\t\n\\end{figure}\n\\par Figure \\ref{7xx} shows the measurement of the virtual photon-proton cross section $\\sigma_{\\gamma^{*}p}^\\text{eff}$ as a function of $Q^2$ corresponding to different values of $W$. The total cross section $\\sigma_{\\gamma^*p}^{\\text{tot}}$ is often expressed as a function of $Q^2$ and invariant mass $W$. The standard relation between $W$, $x$ and $Q^2$ is $W=\\sqrt{Q^2(1-x)\/x}$. Since $Q^2\\approx syx$, for small-$x$ we can have an approximate relationship between $W$ and $y$ i.e. $W^2\\simeq sy$ which we have used in our calculations. HERA measurement for $\\sigma^{\\text{eff}}_{\\gamma^{*}p}$ from H1 \\cite{40} ($\\sqrt{s}= 318\\text{ GeV}$) and ZEUS \\cite{81} ($\\sqrt{s}= 300\\text{ GeV}$) are included for comparison with our theoretical prediction. We have chosen $\\sqrt{s}= 318\\text{ GeV}$ for our calculation analogous to H1 measurement. The precision of the data is such that their errors are hardly visible. Both the HERA data and theoretically measured cross sections for different values of $W$ are multiplied by the factors multiple of 2 as indicated in the figure. For $Q^2\\gtrsim3 \\text{ GeV}^2$, an excellent agreement between the theory and HERA data can be observed for the wide range of $W$, while for $Q^2<3 \\text{ GeV}^2$, the slop of the QCD prediction seems unsatisfactory. This indicates the inadequacy of perturbative QCD at very low $Q^2$ ($<3 \\text{ GeV}^2$) and roughly provides a lower bound of $Q^2$ to our theory.\n\n\n\n\n\n\n\n\n\n\\section{Conclusion}\\label{con}\n\n\\par In conclusion, we have presented a phenomenological study on the behavior of unintegrated gluon distribution at small-$x$ and moderate $k_T^2$ region particularly $10^{-6}\\leq x\\leq 10^{-2}$ and $2 \\text{ GeV}^2\\leq k_T^2\\leq 1000 \\text{ GeV}^2 $ which is also the accessible kinematic range to experiments performed at HERA $ep$ collider. In the beginning, we have improved the MD-BFKL equation supplementing so-called kinematic constraint on it. Then we have solved this unitarized BFKL equation analytically in order to study $x$ and $k_T^2$ dependence of unintegrated gluon distribution function. Our prediction of gluon distribution is contrasted with that of modified BK equation as well as global datasets NNPDF 3.1sx and CT 14. The $x$ evolution of gluon distribution shows the singular $x^{-\\lambda}$ type behavior of gluon evolution tamed by shadowing correction. We found that for intense shadowing $(R= 2 \\text{ GeV}^{-1})$, towards smaller x, certainly from $x=10^{-4}$, the gluon distribution emerges from rapid BFKL growth which indicates the dominance of gluon shadowing. While in case of conventional shadowing $(R= 5 \\text{ GeV}^{-1}) $ an appreciable modification of BFKL behavior can only be seen from $x= 10^{-5}$. The $k_T^2$ dependence of unintegrated gluon distribution is also studied. Although obvious suppression due to net shadowing correction is seen in our studies, however no significant saturation phenomenon is observed. The reason is the nonlinear contribution term is suppressed by the factor $1\/k_T^2$ at large values of $k_T^2$.\n\n\\par We have obtained a more general solution of KC improved MD-BFKL equation implementing a pre-defined input distribution on it. This has allowed us to visualize the gluon evolution in three dimensions $\\mathbb{R}^3$ into $x\\text{-}k_T^2$ phase space. We have also shown the sensitiveness of unintegrated gluon distribution towards $R$ and $\\lambda$ using density plots in $x\\text{-}k_T^2$ plane.\n\n\\par An important achievement of obtaining an analytical solution in this work is its implication on qualitative studies on geometrical scaling which is presented in Sect. \\ref{critical}. Starting from a basic concept of differential geometry and knowledge of level curves we have managed to obtain an equation of the critical boundary which is supposed to separate low and high gluon density regions.\n\\par In Sect. \\ref{hera} we have studied the small-$x$ dependence of the structure function $F_2$ and $F_L$ obtained via $k_T$-factorization formula $F_i=f\\otimes F_i^{(0)}$. Here the unintegrated gluon distribution $f(x,k_T^2)$ is taken from our analytical solution of KC improved MD-BFKL equation setting boundary condition at $x_0=0.01$. Surprisingly the quantitative size of the shadowing correction to $F_2$ and $F_L$ is found to be very small and the structure functions seem to hold the singular $x^{-\\lambda}$ behavior. Even at intense shadowing condition $(R= 2 \\text{ GeV}^{-1})$ the $F_2$ structure function is found to be suppressed only by $10\\%$. In addition we have measured $e^{\\pm}p$ reduced cross section as well as equivalent $\\gamma^*p$ longitudinal to transverse cross section ratio $R(x,Q^2)$ from the knowledge of $F_2$ and $F_L$. Our results are compared with recent high precision HERA measurements. The comparison reveals a good agreement between our theory and DIS data.\n\\par In Sect. \\ref{hera2} we have examined the virtual photon-proton effective cross section, particularly in the transition region from photoproduction to deep inelastic scattering. The quantity $\\sigma_{\\gamma^{*}p}^\\text{eff}$ serves the role of the total cross section if small inelasticity $y$ is concerned. We have compared our predicted $\\sigma_{\\gamma^{*}p}^\\text{eff}$ with the HERA data of two dedicated runs \"NVX'99\" and \"SVX'00\" by H1 as well as ZEUS for the transition region. Our theoretical prediction shows well-consistency with HERA data particularly upto $Q^2\\sim3 \\text{ GeV}^2$ in the transition region.\n\n\n\\par In summary, there are several attractive features of our present study. First, we have able to predict a wide range of physical quantities, starting right from our solution for unintegrated gluon density. Secondly, all analysis is performed in terms of relatively small numbers of parameters. Moreover, the idea developed in this work for studying geometrical scaling can be implemented in any other framework. Finally, a very significant feature of this analysis is that we have considered two extreme possibilities of shadowing which can be distinguished by experimental data. In the end, we conclude that, as per feasibility towards HERA DIS data is concerned, the KC improved MD-BFKL equation could be a reliable framework for exploring high energy physics over a wide range of $x$ and $k_T^2$ which is also relevant for LHC probe and future collider phenomenology. \n\n\n\\section*{Acknowledgments}\nWe thank Prof.~Johannes Bluemlein, DESY (Hamburg, Germany) and Prof.~Dieter Schildknecht, Bielefeld university (Germany) for their valuable comments. Two of us ( P.P. and M.L.) acknowledge Department of Science and Technology (DST), India (grant DST\/INSPIRE Fellowship \/2017 \/IF160770) and Council of Scientific and Industrial Research (CSIR), New Delhi (grant 09\/796(0064)2016-EMR-I) respectively for the financial assistantship. \n\n\n\n\n\\bibliographystyle{spphys}\n","meta":{"redpajama_set_name":"RedPajamaArXiv"}}