id
stringlengths
11
12
text
stringlengths
119
8.37k
image
imagewidth (px)
2.48k
2.55k
230405623/5
Figure4: Patient distribution of genetic subtype (a), lesion type (b), SBR grade (c) and pCR status (d) in the dataset.
230405623/2
Figure 1: Example breast CDI s images from the Cancer-Net BCa open-source benchmark dataset for each SBR grade. a part of a global open-source initiative dedicated to accelerating advancement in machine learning to aid clinicians in the fight against cancer. ## Methodology To construct the Cancer-Net BCa benchmark dataset, we produced CDI s acquisitions for a pre- treatment (T0) patient cohort of 253 patient cases across 10 institutions via the American College of Radiology Imaging Network (ACRIN) 6698/I-SPY2 study [3 – 6]. More specifically, acquisitions were conducted with a four b-value imaging protocol (0 s/mm $^{2}$, 100 s/mm $^{2}$, 600 s/mm $^{2}$, 800 s/mm $^{2}$, 3-direction) on a 1.5 or 3.0 Tesla scanner using a dedicated breast radiofrequency coil. The pixel spacing for the acquisitions ranged from 0.83 mm to 2.08 mm with a median of 1.29 mm, with both slice thickness and spacing between slices ranged from 4.0 to 5.0 mm with a median of 4.0. The native and synthetic signals produced via a signal synthesizer were mixed together to obtain a final CDI s signal [1]. Each patient case is also associated with one of three possible SBR grades: I (Low), II (Intermediate), and III (High). Example images from each SBR type is shown in Fig. 1. The pCR state after neoadjuvant chemotherapy (No pCR/pCR) is also provided for each patient, with an example of each pCR state shown in Fig. 2. ## Results and Discussion
230405623/1
# A Multi-Institutional Open-Source Benchmark Dataset for Breast Cancer Clinical Decision Support using Synthetic Correlated Diffusion Imaging Data Chi-en Amy Tai Hayden Gunraj Alexander Wong Vision and Image Processing Lab, University of Waterloo {amy.tai, hayden.gunraj, alexander.wong} @uwaterloo.ca Recently, a new form of magnetic resonance imaging (MRI) called synthetic correlated diffusion (CDI s) imaging was introduced and showed considerable promise for clinical decision support for cancers such as prostate cancer when compared to current gold-standard MRI techniques. However, the efficacy for CDI s for other forms of cancers such as breast cancer has not been as well- explored nor have CDI s data been previously made publicly available. Motivated to advance efforts in the development of computer-aided clinical decision support for breast cancer using CDI s, we introduce Cancer-Net BCa, a multi-institutional open-source benchmark dataset of volumetric CDI s imaging data of breast cancer patients. Cancer-Net BCa contains CDI s volumetric images from a pre-treatment cohort of 253 patients across ten institutions, along with detailed annotation meta- data (the lesion type, genetic subtype, longest diameter on the MRI (MRLD), the Scarff-Bloom-Richardson (SBR) grade, and the post-treatment breast cancer pathologic complete response (pCR) to neoadjuvant chemotherapy). We further examine the demographic and tumour diversity of the Cancer-Net BCa dataset to gain deeper insights into potential biases. Cancer-Net BCa is publicly available as a part of a global open-source initiative dedicated to accelerating advancement in machine learning to aid clinicians in the fight against cancer. ## Introduction A new form of magnetic resonance imaging (MRI) called synthetic correlated diffusion (CDI s) imaging was recently introduced and showed considered promise for clinical decision support for cancers such as prostate cancer when compared to current gold-standard MRI techniques such as T2-weighted (T2w) imaging, diffusion-weighted imaging (DWI), and dynamic contrast-enhanced (DCE) imaging [1]. However, the efficacy for CDI s for other forms of cancer such as breast cancer has not been as well-explored nor have CDI s data been previously made publicly available. The development of computer-aided clinical decision support for breast cancer using CDI s has begun to be analyzed and shown to have superior results compared to other gold-standard imaging for the prediction of breast cancer patient response from neoadjuvant chemotherapy prior to treatment [2]. Motivated to advance efforts in the development of computer-aided clinical decision support for breast cancer using CDI s for diagnosis, prognosis/grading, treatment planning and more, we in- troduce Cancer-Net BCa, a multi-institutional open-source benchmark dataset of volumetric CDI s imaging data of breast cancer patients with detailed annotation metadata for each patient. We further examine the demographic and grade diversity of the Cancer-Net BCa dataset to gain deeper insights into potential biases. The Cancer-Net BCa benchmark dataset has been made publicly available 1 as 1 https://www.kaggle.com/datasets/amytai/cancernet-bca
170404598/16
SIMONA NISTOR where $$ \stackrel{\mathclap{\tiny\mbox{$N$}}}{\Riem}\left(X_1,X_2\right) =R^N\left(X_1,X_2,X_1,X_2\right). $$ It is clear that $\langle H,\eta_\alpha\rangle=0$, and then $\trace A_{\eta_\alpha}=2\langle H,\eta_\alpha\rangle=0$, for any $\alpha\in \left\{1,2,\cdots n-3\right\}$. As $A_{\eta_\alpha}$ is symmetric, we note that $\det A_{\eta_\alpha}\neq 0$ for any $\alpha$ and $\sum_{\alpha=1}^{n-3}\det A_{\eta_\alpha}\leq 0$. Then, we get $$ K-\stackrel{\mathclap{\tiny\mbox{$N$}}}{\Riem}\left(X_1,X_2\right)\leq \det A_\eta. $$ Let us consider $\mu_1$ and $\mu_2$ the principal curvatures of $A_\eta$. Then \begin{align*} \det A_\eta =& \mu_1\mu_2=\frac{\left(\mu_1+\mu_2\right)^2-\left(\mu_1^2+\mu_2^2\right)}{2} \\ = & \frac{\left(\trace A_\eta\right)^2-\left|A_\eta\right|^2}{2} \\ = & \frac{4|H|^2-\left|A_\eta\right|^2}{2}. \end{align*} From $\eqref{ineq-1}$ one obtains $$ \left|A_\eta\right|^2\leq 4|H|^2-2K+2\stackrel{\mathclap{\tiny\mbox{$N$}}}{\Riem}\left(X_1,X_2\right). $$ Since $K\geq 0$ and $\stackrel{\mathclap{\tiny\mbox{$N$}}}{\Riem}\ \leq k_0$, it follows that $$ \left|A_\eta\right|^2\leq 4|H|^2+2k_0. $$ Therefore, $\left|A_\eta\right|^2$ is bounded from above by the constant $4|H|^2+2k_0$ and then $\left|A_H\right|^2$ is bounded from above. It is well known that a complete surface with $K\geq 0$ is parabolic (see [11]), i.e., any subharmonic function bounded from above is constant. Thus, as $\left|S_2\right|^2$ is bounded from above and subharmonic, it follows that $\left|S_2\right|^2$ is a constant. Using $\left|\nabla S_2\right|^2=16\left|\nabla A_H\right|^2$ and $\eqref{Laplac-S2}$ one obtains that $\nabla A_H=0$, and therefore $M$ is flat or pseudoumbilical. 5.1. ## Exemples of submanifolds with $\nabla A_H=0$ . As we have seen, a $PMC$ surface in a space form $N^n(c)$, $n\geq 4$, is trivially biconservative. But, if the surface is only $CMC$ then it is not necessarily biconservative. In [17] it was proved that if a surface is biconservative and $CMC$ in $N^4(c)$, with $c\neq0$, then the surface has to be $PMC$, i.e., the trivial case for our problem. We just recall here that, if $c=1$, then a $PMC$ surface in $\mathbb{S}^4$ is either a minimal surface of a small hypersphere of radius $a$, $a\in (0,1)$, in $\mathbb{S}^4$, or a $CMC$ surface in a small or great hypersphere in $\mathbb{S}^4$ (see [24,25]). Of course, if we consider a $CMC$ biconservative surface $M^2$ of genus $0$ in $\mathbb{R}^4$, it is pseudoumbilical and therefore it is $PMC$, i.e., $M^2$ is a $2$ -sphere (see [10]). In $\mathbb{R}^4$, there were obtained all $CMC$ biconservative surfaces which are not $PMC$. They are given by the isometric immersion $\varphi:\mathbb{R}^2\rightarrow\mathbb{R}^4$ defined by $$ \varphi(u,v)=\overline{\gamma}(u)+(v+a)\overline{e}_4, $$ where $\overline{\gamma}:\mathbb{R}\rightarrow\mathbb{R}^3$ is a smooth curve parametrized by arc length with positive constant curvature $k$ and free torsion $\tau$. By direct computation we obtain that the second fundamental form of the surface is given by $$ B\left(\partial_u,\partial_u\right)=k(u)N(u), \qquad B\left(\partial_u,\partial_v\right)=0, \qquad B\left(\partial_v,\partial_v\right)=0, $$
170404598/4
SIMONA NISTOR - (1) the stress-bienergy tensor of $\varphi$ is determined by $$ S_2= - \frac{m^2}{2}|H|^2 I + 2m A_H; $$ $\trace S_2 = m^2|H|^2\left(2-\frac{m}{2}\right)$; - (3) the relation between the divergence of $S_2$ and the divergence of $A_H$ is given by $$ \Div S_2=-\frac{m^2}{2}\grad \left(|H|^2\right)+2m\Div A_H; $$ $\left|S_2\right|^2=m^4|H|^4\left(\frac{m}{4}-2\right)+4m^2\left|A_H\right|^2$. Remark 2.8. From equation $\eqref{divergente0}$, we see that if $M$ is biconservative it does not follow that $\Div A_H$ automatically vanishes. In fact, only when $|H|$ is constant the biconservativity is equivalent to $\Div A_H=0$. ## 3. Other characterizations of biconservative submanifolds In this section we will characterize the biconservative submanifolds which satisfy some additional geometric hypotheses. We begin with a study on the basic properties of submanifolds with $A_H$ parallel, as they are the “ simplest ” biconservative surfaces. First, we define the principal curvatures of a submanifold $M^m$ of $N^n$ as being the eigenvalue functions of $A_H$. Proposition Let $\varphi:M^m\rightarrow N^n$ be a submanifold and $\lambda_1\geq \cdots \geq \lambda_m$ the principal curvatures of $M$. If $\nabla A_H=0$, then: $M$ is - biconservative; $\lambda_i$ are constant functions on $M$, in particular $M$ is $CMC$; $A_{\nabla^\perp_X H}(Y)-A_{\nabla^\perp_Y H}(X)=\left(R^N(X,Y)H\right)^T$, for any $X,Y\in C(TM)$; $\trace A_{\nabla^\perp_\cdot H}(\cdot)=-\trace\left(R^N(\cdot,H)\cdot\right)^T$. Proof. For the sake of completeness, we give only the proof of the second item. More precisely, we show that $\lambda_i$ are constant functions on $M$. Let us consider an arbitrary point $p\in M$. Since $A_H(p)$ is symmetric, then $A_H(p)$ is diagonalizable. We denote by $\lambda_{1,p}\geq\cdots\geq\lambda_{m,p}$ the eigenvalues of $A_H(p)$ and then we define the continuous functions $\lambda_i:M\rightarrow \mathbb{R}$, $\lambda_i(p)=\lambda_{i,p}$, for any $p\in M$ and any $i=\overline{1,m}$. Further, we consider $\left\{e_i\right\}_{i=\overline{1,m}}$ an orthonormal basis in $T_p M$ which diagonalize $A_H(p)$, i.e., $\left(A_H(p)\right)\left(e_i\right)=\lambda_i(p)e_i$, for any $i=\overline{1,m}$. Consider $q\in M$, $q\neq p$, and $\gamma:[a,b]\rightarrow M$ a smooth curve such that $\gamma(a)=p$ and $\gamma(b)=q$. We define the vector fields $E_i=E_i(t)$, along $\gamma$, such that $DE_i(t)/dt=0$, for any $t$ and $E_i(a)=e_i$. It is easy to see that $W(t)=\left(A_H(\gamma(t))\right)\left(E_i(t)\right)$ is also a vector field along $\gamma$ and \begin{align*} \frac{DW}{dt}(t)= & \left(\nabla_{\gamma'(t)}A_H\right)(E_i(t))+A_H\left(\frac{D E_i}{dt}(t)\right)\\ = & \ 0. \end{align*} Now, since $D\left(\lambda_i(p)E_i\right)(t)/dt=0$, we get that $W(t)$ and $\lambda_i(p)E_i$ are parallel vector fields along $\gamma$. Since for $t=a$ they are equal, it follows that they coincide for any $t$, and in particular, for $t=b$. Therefore, $\lambda_i(p)$, $i=\overline{1,m}$, are eigenvalues of $A_H(q)$. As $q$ was chosen in an arbitrary way, we get that $\lambda_i$ are constant functions on $M$, for any $i=\overline{1,m}$.
170404598/1
# ON BICONSERVATIVE SURFACES SIMONA NISTOR Abstract. We study in a uniform manner the properties of biconservative sur- faces in arbitrary Riemannian manifolds. Biconservative surfaces being charac- terized by the vanishing of the divergence of a symmetric tensor field $S_2$ of type $(1,1)$, their properties will follow from general properties of a symmetric tensor field of type $(1,1)$ with free divergence. We find the link between the biconser- vativity, the property of the shape operator $A_H$ to be a Codazzi tensor field, the holomorphicity of a generalized Hopf function and the quality of the surface to have constant mean curvature. Then we determine the Simons type formula for biconservative surfaces and use it to study their geometry. ## 1. Introduction In the last decade the theory of biconservative submanifolds proved to be a very interesting research topic (see, for example, [1,4–7,16,19,22,23]). This theory arose from the theory of biharmonic submanifolds, but the class of biconservative subman- ifolds is richer than the later one. Let $\left(M^m,g\right)$ and $\left(N^n,h\right)$ be two Riemannian manifolds. A biharmonic map is a critical point of the bienergy functional $$ E_2:C^{\infty}(M,N)\rightarrow\mathbb{R},\quad E_{2}(\varphi)=\frac{1}{2}\int_{M}|\tau(\varphi)|^{2}\ v_g, $$ where $\tau(\varphi)$ is the tension field of a smooth map $\varphi:M\rightarrow N$, and it is characterized by the vanishing of the bitension field $\tau_2(\varphi)$ (see). If $\varphi:\left(M^m,g\right)\rightarrow \left(N^n,h\right)$ is a biharmonic map and a Riemannian immersion, then $M$ is called a biharmonic submanifold of $N$. According to D. Hilbert (see [9]), to a functional $E$ we can associate a symmetric tensor field $S$ of type $(1,1)$, called the stress-energy tensor, which is conservative, i.e., $\Div S=0$, at the critical points of $E$. In the particular case of the bienergy functional $E_2$, G. Y. Jiang (see [13]) defined the stress-bienergy tensor $S_2$ by \begin{align*} \langle S_2(X),Y\rangle=&\frac{1}{2}|\tau(\varphi)|^2\langle X,Y\rangle+\langle d\varphi,\nabla\tau(\varphi)\rangle\langle X,Y\rangle\ &-\langle d\varphi(X),\nabla_Y\tau(\varphi)\rangle-\langle d\varphi(Y),\nabla_X\tau(\varphi)\rangle, \end{align*} and proved that $$ \Div S_2=-\langle\tau_2(\varphi),d\varphi\rangle. $$ Therefore, if $\varphi$ is biharmonic, then $\Div S_2=0$ (see [13,14]). One can see that if $\varphi:\left(M^m,g\right)\rightarrow\left(N^n,h\right)$ is a Riemannian immersion then $\Div S_2=0$ if and only if the tangent part of the bitension field vanishes. A subman- ifold $M$ is called biconservative if $\Div S_2=0$. Key words and phrases. Biconservative surfaces, biharmonic submanifolds, mean curvature vec- tor field, Codazzi tensor fields. The author was supported by a grant of the Romanian National Authority for Scientific Research and Innovation, CNCS - UEFISCDI, project number PN-II-RU-TE-2014-4-0004.
170404598/3
where $B\in C(\odot^2 T^*M\otimes NM)$ is the second fundamental form of $M$ in $N$ and the $\trace$ is considered with respect to the metric on $M$. ## 2. Preliminaries First we recall some notions, formulas and general results about tensor fields and submanifolds that we will use later. It is well-known that a symmetric tensor field $T$ of type $(1,1)$ on a Riemannian manifold $\left(M^m,g\right)$ can be identified with a symmetric tensor field $\tilde{T}$ of type $(0,2)$ $$ \langle T(X),Y\rangle=\tilde{T}(X,Y), \qquad X,Y\in C(TM), $$ and, we will use the same notation $T$ instead of $\tilde{T}$. Proposition 2.1. Let $\left(M^m,g\right)$ be a Riemannian manifold and consider $T$ and $S$ two symmetric tensor fields of type $(1,1)$. Then $$ \langle\Delta^R T,S\rangle = \langle \nabla T,\nabla S \rangle - \Div Z, $$ with $\Delta^R T = -\trace \nabla^2 T$, $Z\in C(TM)$, $Z=\langle \nabla_{X_i} T,S\rangle X_i$, where $\left\{X_i\right\}_{i=\overline{1,m}}$ is an orthonormal local frame field. Proposition 2.2. Let $\left(M^m,g\right)$ be a Riemannian manifold and consider $T$ a sym- metric tensor field of type $(1,1)$ and $\alpha$ is a smooth function on $M$. Then $$ \Div \left(T\left(\grad \alpha\right)\right) = \langle \Div T,\grad \alpha \rangle + \langle T, \Hess \alpha\rangle, $$ Definition A submanifold $\varphi:M^m\rightarrow N^n$ is called pseudoumbilical if $A_H=|H|^2 I$, where $I$ is the identity tensor field of type $(1,1)$. Using the Codazzi equation, we easily find the next result. Proposition Let $\varphi:M^m\rightarrow N^n$ be a submanifold. Then $$ \trace \nabla A_H = \frac{m}{2}\grad \left(|H|^2 \right) + \trace A_{\nabla_{\cdot}^\perp H}(\cdot)+ \trace\left(R^N(\cdot,H)\cdot\right)^T. $$ Corollary Let $\varphi:M^m\rightarrow N^n(c)$ be a submanifold, $c\in\mathbb{R}$. Then $$ \trace \nabla A_H = \frac{m}{2}\grad \left(|H|^2 \right) + \trace A_{\nabla_{\cdot}^\perp H}(\cdot). $$ Let $\varphi:M^m\rightarrow N^n$ be a submanifold. Computing $\tau_2(\varphi)$ by splitting it in the tangent and in the normal part and using (2.3) we get the following characterizations for biconservative submanifolds (various expressions for $\tau_2(\varphi)$ were obtained in [2, 14,20,21]). Proposition 2.6. Let $\varphi:M^m\rightarrow N^n$ be a submanifold. Then the following condi- tions are equivalent: $M$ is - biconservative; $\trace A_{\nabla^\perp_{\cdot} H}(\cdot)+\trace \nabla A_H +\trace \left(R^N(\cdot,H)\cdot\right)^T=0$; $\frac{m}{2}\grad\left(|H|^2\right)+2\trace A_{\nabla^\perp_{\cdot} H}(\cdot) + 2\trace \left(R^N(\cdot,H)\cdot\right)^T=0$; $2\trace \nabla A_H-\frac{m}{2}\grad\left(|H|^2\right)=0$. We end this section with the following result. Proposition 2.7. Let $\varphi:M^m\rightarrow N^n$ be a submanifold. Then we have:
201001150/5
Figure3: Correlation between different similarity mea- sures and diversity measure and the improvement ($\Delta$) due to domain-specific BERT models. ## Analysis After we empirically show the importance of se- lecting in-domain source data, the next question is: can we find a cost-effective way to nominate in-domain source data? ## Measuring Similarity We use three measures of the similarity between source and target data. We then observe whether these similarity values correlate with the usefulness of pretrained models in $\S$ 5.2. Language model perplexity (PPL) has been used to provide a proxy to estimate corpus sim- ilarity (Baldwinetal., 2013). We construct Kneser- Ney smoothed 3-gram models (Heafield, 2011) on source data and use the perplexity of target data relative to these language models as the similarity between source and target data. Jensen-Shannon divergence (JSD), based on term distributions, has been successfully used for domain adaptation (Ruder and Plank, 2017). We first measure the probability of each term (up to 3-gram) in source and target data, separately. Then, we use the Jensen-Shannon divergence between these two probability distributions as the similarity between source and target data. Targetvocabularycovered(TVC) measures the percentage of the target vocabulary present in the source data, where only content words (nouns, verbs, adjectives) are counted. Dai et al. (2019) show that it is very informative in predicting the effectiveness of pretrained word vectors. In addition, RuderandPlank (2017) show that the diversity of source data is as important as do- main similarity for domain adaptation. Inspired by this, we also explore a very simple diversity mea- sure: type token ratio (TTR, $\frac{\# \, \text{unique} \, \text{tokens}}{\# \, \text{tokens}}$), that measures the lexical diversity of the source data. To mitigate the impact of source data size on these measurements, for each source data, we sam- ple five sub-corpora, each of which contains 10M tokens. Then we measure the similarity of source and target data and the diversity of source data as the average values of these sub-corpora. ## Correlation Analysis To analyze how the effectiveness of domain- specific BERT models correlate to the similarity be- tween source and target data, we employ the Pear- son correlation analysis to find out the relationships between improvements due to domain-specific BERT models and similarity between source and target data. For example, considering the BTC task, we use the performance of the original BERT as baseline, and measure the improvement due to Twitter BERT as $1.0$, whereas the corresponding value using BioBERT is $-2.9$. Note that we repeat all the experiments five times; therefore, we collect 300 source-target data points in total. The correlation results are visualized in Figure 3. JSD has the strongest correlation (0.519) with the improvement due to domain-specific models, while the other two measures also have modest correla- tion ($0.481$ for PPL and $0.436$ for TVC). Recall that the calculation of JSD takes uni-grams, bi- grams and tri-grams into consideration, whereas PPL considers tri-grams only and the TVC consid- ers uni-grams only. Correlations between different measures indicate that these measures are able to reach agreement on whether source and target are similar. We find no correlation between the TTR of source data and the improvement. ## Summary We conduct a case study of pretraining BERT on social media text. Through extensive experiments, we show the importance of selecting in-domain source data. Based on empirical analysis, we rec- ommend measures to help select pretraining data for best performance on new applications. ## Acknowledgments We would like to thank anonymous reviewers for their helpful comments. XD also thanks Shubin Du and Ying Zhou for early investigation of this work. XD is supported by Sydney University’s En- gineering and Information Technologies Research Scholarship and a CSIRO Data61 top up scholar- ship.
201001150/4
\begin{tabular}{c|l|c|c|c|c|c|c} \rightarrowprule %&&\multicolumn{6}{c}{\bf BERT}\bf Target Text type& \bf Corpus & \bf BERT & \bf Bio & \bf Clinical & \bf Sci & \bf Twitter & \bf Forum & & (3.3B) & (18B) & (0.5B) & (3.1B) & (0.9B) & (0.6B) \midrule \multirow{4}{*}{Tweets} & Airline ($C$) & 80.5\scriptsize{$\pm$ 0.3} & 79.0\scriptsize{$\pm$ 0.5} & 78.8\scriptsize{$\pm$ 0.8} & 78.8\scriptsize{$\pm$ 0.9} & 80.8\scriptsize{$\pm$ 0.6} & \underline{\bf 81.6\scriptsize{$\pm$ 0.5}} % Dec-4 & BTC ($N$) & 78.0\scriptsize{$\pm$ 0.5} & 75.2\scriptsize{$\pm$ 0.3} & 76.9\scriptsize{$\pm$ 0.5} & 77.4\scriptsize{$\pm$ 0.4} & \bf 79.0\scriptsize{$\pm$ 0.5} & 77.0\scriptsize{$\pm$ 0.4} % Dec-4 & SMM4H-18 task3 ($C$) & 76.5\scriptsize{$\pm$ 0.9} & 75.4\scriptsize{$\pm$ 1.1} & 75.6\scriptsize{$\pm$ 0.7} & 75.4\scriptsize{$\pm$ 1.0} & 77.0\scriptsize{$\pm$ 1.0} & \bf 77.2\scriptsize{$\pm$ 1.3} % Task3 % Dec-4 & SMM4H-18 task4 ($C$) & 89.4\scriptsize{$\pm$ 0.5} & 87.7\scriptsize{$\pm$ 0.4} & 88.1\scriptsize{$\pm$ 0.8} & 88.7\scriptsize{$\pm$ 0.8} & 90.3\scriptsize{$\pm$ 0.3} & \underline{\bf 91.1\scriptsize{$\pm$ 0.6}} % Task4 % Dec-4 %& WNUT-16 ($N$) & 49.7 & 38.5 & 43.9 & 46.2 & \bf 49.9 & 46.7 % Nov-30 \hline \multirow{4}{*}{Forum} & CADEC ($N$) & 71.9\scriptsize{$\pm$ 0.6} & 72.1\scriptsize{$\pm$ 0.6} & 72.1\scriptsize{$\pm$ 0.8} & \underline{\bf 73.2\scriptsize{$\pm$ 0.4}} & 72.1\scriptsize{$\pm$ 1.0} & 72.9\scriptsize{$\pm$ 0.6} % Nov-30 % & CADEC ($C$) & 91.6 & \bf 92.2 & 91.4 & 92.1 & 90.9 & 91.9 % & SMM4H-18 T3 ($C$) & 76.3 & 76.4 & 76.4 & 74.9 & \bf 77.7 & 77.5 & SemEval-14 laptop ($N$) & 81.1\scriptsize{$\pm$ 0.8} & 79.3\scriptsize{$\pm$ 0.3} & 78.5\scriptsize{$\pm$ 0.4} & \bf 81.6\scriptsize{$\pm$ 1.1} & 81.3\scriptsize{$\pm$ 0.6} & 81.4\scriptsize{$\pm$ 1.1} % Nov-30 & SemEval-14 restaurant ($N$) & 87.5\scriptsize{$\pm$ 0.6} & 84.9\scriptsize{$\pm$ 0.3} & 85.5\scriptsize{$\pm$ 0.7} & 86.7\scriptsize{$\pm$ 0.5} & 87.4\scriptsize{$\pm$ 0.7} & \underline{\bf 89.3\scriptsize{$\pm$ 0.5}} % Dec-4 & SST-2 ($C$) & 92.4\scriptsize{$\pm$ 0.2} & 91.1\scriptsize{$\pm$ 0.5} & 90.4\scriptsize{$\pm$ 0.3} & 91.4\scriptsize{$\pm$ 0.4} & 92.3\scriptsize{$\pm$ 0.4} & \underline{\bf 93.4\scriptsize{$\pm$ 0.4}} % Dec-2 %& & Amazon review & 61.5 & 60.5 & 60.4 & 60.2 & 61.3 & \bf 62.6 \hline \multirow{3}{*}{Non-social media} & EBM ($N$) & 41.5\scriptsize{$\pm$ 0.5} & 42.1\scriptsize{$\pm$ 0.2} & 41.1\scriptsize{$\pm$ 0.5} & \bf 42.4\scriptsize{$\pm$ 0.7} & 40.5\scriptsize{$\pm$ 0.5} & 41.5\scriptsize{$\pm$ 0.5} % Nov-30 & i2b2-10 ($N$) & 85.8\scriptsize{$\pm$ 0.1} & \underline{\bf 87.4\scriptsize{$\pm$ 0.2}} & \underline{\bf 87.4\scriptsize{$\pm$ 0.1}} & 87.3\scriptsize{$\pm$ 0.2} & 84.8\scriptsize{$\pm$ 0.2} & 85.2\scriptsize{$\pm$ 0.1} % Nov-30 & JNLPBA ($N$) & 72.5\scriptsize{$\pm$ 0.3} & \underline{\bf 74.2\scriptsize{$\pm$ 0.2}} & 71.9\scriptsize{$\pm$ 0.1} & 73.6\scriptsize{$\pm$ 0.3} & 72.2\scriptsize{$\pm$ 0.2} & 72.5\scriptsize{$\pm$ 0.2} % Nov-30 & Paper Field ($C$) & 74.5\scriptsize{$\pm$ 0.1} & 74.3\scriptsize{$\pm$ 0.1} & 73.3\scriptsize{$\pm$ 0.1} & \underline{\bf 75.1\scriptsize{$\pm$ 0.1}} & 74.1\scriptsize{$\pm$ 0.1} & 73.3\scriptsize{$\pm$ 0.2} % & & CoNLL03 & \bf 91.0 & 89.6 & 90.5 & 87.5 & 90.1 & 88.4 \bottomrule \end{tabular} Table2: Effectiveness of different BERT models, evaluated on downstream tasks. # tokens in each pretraining data are listed in brackets. $C$: Classification task, for which we report macro-F1; $N$: NER task, for which we report span-level micro-F1. We repeat all experiments five times with different random seeds. Mean values are reported. underline: the best result is significantly better than the second best result (paired student’s t-test, p: $0.05$). (a) False positives. (b) False negatives. Figure2: Error predictions on CADEC. Dotted line cir- cle: errors by the BERT model. Dashed line: YelpBERT. Solid line: SciBERT. representations on short tweets. We also observe that, when domain-specific models are applied on a target task with out-of- domain data, they achieve much lower results than the original BERT. For example, BioBERT achieves lower results than the original BERT on 7 out of 8 target social media tasks. It only achieves a better result on CADEC, which is about medica- tions. Recall that all these domain-specific BERT models use the pretrained weights of the original BERT as initialization. On one hand, we argue that this observation may challenge the conven- tional wisdom that the larger the pretraining data is, the better the pretrained model is. Training on out-of-domain source data may cause negative im- pact, at least for the two-stage pretraining approach we consider. On the other hand, this observation reinforces recent work showing the importance of task-adaptive pretraining (Gururanganetal., 2020). ErroranalysisonCADEC We conduct an error analysis on CADEC, because it is at the intersec- tion between social media tenor (online posts) and medication field (adverse drug events), and thus could be similar to multiple sources. We compare the error predictions by the two best performing BERT models – ForumBERT and SciBERT, as well as the baseline BERT model. In Figure 2a, we observe that both domain-specific BERT models can reduce greatly the number of false positives made by the baseline BERT. Specifically, 159 false positives made by the baseline BERT are fixed by the domain-specific BERT models. However, domain-specific BERT models do not reduce a lot the number of false negatives – gold mentions not recognized. There are 258 gold mentions recog- nized by none of three models, and only 41 false negatives by the baseline BERT are fixed by the domain-specific BERT models (Figure 2b).
201001150/1
# Cost-effective Selection of Pretraining Data: A Case Study of Pretraining BERT on Social Media \begin{tabular}{cccc} Xiang Dai$^{1,2}$ & Sarvnaz Karimi$^{1}$ & \textbf{Ben Hachey$^{3}$} & \textbf{Cecile Paris$^{1}$} \end{tabular} \begin{tabular}{cccc} \multicolumn{4}{c}{$^{1}$CSIRO Data61, Sydney, Australia}\multicolumn{4}{c}{$^{2}$University of Sydney, Sydney, Australia}\multicolumn{4}{c}{$^{3}$Harrison ai, Sydney, Australia}\multicolumn{4}{c}{\tt \{dai.dai,sarvnaz.karimi,cecile.paris\}@csiro.au}\multicolumn{4}{c}{\tt [email protected]} \end{tabular} Recent studies on domain-specific BERT mod- els show that effectiveness on downstream tasks can be improved when models are pre- trained on in-domain data. Often, the pretrain- ing data used in these models are selected based on their subject matter, e.g., biology or com- puter science. Given the range of applications using social media text, and its unique language variety, we pretrain two models on tweets and forum text respectively, and empirically demon- strate the effectiveness of these two resources. In addition, we investigate how similarity mea- sures can be used to nominate in-domain pre- training data. We publicly release our pre- trained models at https://bit.ly/35RpTf0. ## Introduction Sequence transfer learning (Ruder, 2019), that pre- trains language representations on unlabeled text (source) and then adapts these representations to a supervised task (target), has demonstrated its effectiveness on a range of NLP tasks (Radford et al., 2018; Devlin et al., 2019; Liu et al., 2019). Approaches vary in model, pretraining objective, pretraining data and adaptation strategy. We con- sider a widely used method, BERT (Devlinetal., 2019). It pretrains a transformer-based model us- ing a masked language model objective and then fine-tunes the model on the target task. We inves- tigate the impact of the domain (i.e., the similar- ity between the underlying distribution of source and target data) of pretraining data on the effec- tiveness of pretrained models. We also propose a cost-effective way to select pretraining data. Recent studies on domain-specific BERT mod- els, which are pretrained on specialty source data, empirically show that, when in-domain data is used for pretraining, target task performance can be im- proved (Lee et al., 2019; Alsentzer et al., 2019; Huang et al., 2019; Beltagy et al., 2019). These publicly available domain-specific BERT models are valuable to the NLP community. However, the selection of in-domain data usually resorts to intu- ition, which varies across NLP practitioners (Dai et al., 2019). According to Halliday and Hasan (1989), the context specific usage of language is affected by three factors: field (the subject matter being discussed), tenor (the relationship between the participants in the discourse and their purpose) and mode (communication medium, e.g., ‘spoken’ or ‘written’). 1 Generally, the selection of pretrain- ing data in existing domain-specific BERT models is based on the field rather than the tenor. For example, BioBERT (Lee et al., 2019) and SciB- ERT (Beltagyetal., 2019) are both pretrained on scholar articles, but on different fields (biology and computer science). We conduct a case study of pretraining BERT on social media text which has very different tenor from existing domain-specific BERT models. Our contributions are two-fold: (1) We release two pre- trained BERT models trained on tweets and forum text, and we demonstrate the effectiveness of these two resources on a range of NLP data sets using social media text; and, (2) we investigate the corre- lation of source-target similarity and task accuracy using different domain-specific BERT models. We find that simple similarity measures can be used to nominate in-domain pretraining data (Figure 1). ## Related Work SelectingdatatopretrainBERT There are two known strategies: (1) collecting very large generic data, such as web crawl and news (Radfordetal., 2019; Liuetal., 2019; Baevskietal., 2019); and, (2) selecting in-domain data, which we refer to as 1 We do not explicitly consider mode in this study, because all data used are written text.
201001150/3
with sequence-pair training (Liuetal., 2019). Twitter We use English tweets ranging from Sep 1 to Oct 30, 2018 4 to pretrain our Twitter BERT. There are in total 60 million English tweets, con- sisting of $0.9$ B tokens. Although we aim to avoid tailored pre-processing strategies to make a fair comparison with other domain-specific BERT mod- els, we find 44% of these tweets contain url and 78% contain other user names (@, if a tweet replies another tweet, @ is added automatically). We thus employ minimal processing by: (1) replacing to- kens starting with ‘@’, referring to a Twitter user’s account name, with a special token [TwitterUser]; and, (2) replacing urls as a special token [URL]. We hypothesize that the surface form of these tokens do not contain useful information. Forum We use local businesses reviews released by Yelp 5 to pretrain our Forum BERT. There are in total five million reviews, consisting of $0.6$ B tokens. No preprocessing is conducted on the text. We used four Nvidia P100 GPUs for the pretrain- ing. Training of each model took seven days. ## Effectiveness of Pretrained BERT Models To evaluate the effectiveness of our pretrained BERT models, we experiment on a range of clas- sification and Named Entity Recognition (NER) data sets. Both text classification and NER are fundamental NLP tasks that can employ generic architectures on top of BERT. For the classifica- tion task, the representation of the first token (i.e., [CLS]) is fed into the output layer for the final prediction. For the NER task, the representations of the first sub-token within each token are taken as input to a token-level classifier to predict the token’s tag. We did not explore more complex ar- chitectures, such as adding LSTM or CRF on top of BERT (Beltagyetal., 2019; Baevskietal., 2019), because our aim is to demonstrate the efficacy of domain-specific BERT models and to observe the impact of pretraining data, rather than to achieve state-of-the-art performance on these data sets. Our BERT results follow the standard two- stage approach of finetuning the pretrained model. Domain-specific BERTs add a stage in the mid- dle: finetuning BERT on domain-specific unlabeled data (cf. Figure 1). ## Target Tasks We use eight target tasks with their text sampled from Twitter and forums, to examine whether our BERT models can lead to improvements, com- pared to the original BERT. These tasks are Air- line 6: classifying sentiment on tweets about ma- jor U.S. airlines; BTC: identifying location, per- son, and organization on tweets (Derczynskietal., 2016); SMM4H-18: classifying whether the user reports an adverse drug events (task3) (Weis- senbacheretal., 2018), or intends to receive a sea- sonal influenza vaccine (task4) on tweets about health (Joshi et al., 2018); CADEC: identifying adverse drug events etc. on reviews about medica- tions (Karimietal., 2015); SemEval-14: identify- ing product or service attributes on reviews about laptops and restaurants (Pontikietal., 2014); SST: classifying sentiment on movie reviews (Socher etal., 2013). In addition, we use four tasks that do not use so- cial media text to investigate how our BERT mod- els perform on out-of-domain target tasks: Paper Field: classifying the research topic based on the title of scholar articles about various fields (Belt- agy et al., 2019); EBM: identifying intervention, outcome etc. on scholar articles about clinical tri- als (Nye et al., 2018); i2b2-10: identifying treat- ment, test and problem on clinical notes about health (Uzuneretal., 2011); JNLPBA: identifying RNA, DNA etc. on scholar articles about biol- ogy (Kimetal., 2004). ## Results We observe that our BERT models achieve the high- est F1 score on 6 out of 8 target tasks that use social media text (Table 2). On CADEC (medica- tions) and SemEval-14 laptop, SciBERT achieves the highest score due to the overlapping fields (i.e., medication and computer hardware, respectively). We note, however, that our Forum BERT achieves very close results. This demonstrates the effec- tiveness of our pretrained models on target tasks using social media text. To our surprise, on target tasks using tweets, forum BERT achieves better results than Twitter BERT on 3 classification tasks. On one hand, this may be explained by Baldwin et al. (2013) ’s observation that forum text is the ‘median’ data, which is similar to all other types of social media text. On the other hand, it also reveals the challenge of pretraining contextual language 4 Internetarchive, Accessed 1 June 2020. 5 YelpChallenge, Accessed 1 June 2020.
230600409/2
common visual prompting. This strategy has been recently attempted in some recent advances, such as FewVLM (Jin et al. 2021) and BLIP-2 (Li et al. 2023). In this paper, we also conduct some toy experiments on VL benchmarks, of which results show that this strategy can help BERT (De- vlinetal.2018) obtain decent capabilities on VL tasks, e.g. 68.0% on VQA2.0 (Goyaletal.2017) and 76.3% on SNLI- VE (Xie et al. 2019). Although the performance still lags behind large-scale vision-language pre-trained (VLP) mod- els, such as CLIP-ViL (Shenetal.2021) and METER (Dou etal.2022), it yields an affordable way for quick VL adap- tion. Nevertheless, visual prompting greatly increases the com- putation complexity of PLMs. For instance, directly adding all the visual patches of ViT (Dosovitskiy et al. 2020) to BERT (Devlin et al. 2018) will increase the already high computation by up to 384%. Moreover, the use of complete image features is often redundant to the VL models, as re- vealed in the previous VL research (Yangetal.2016;Zhou etal.2019,2021). During experiments, we also observe that the deployment of visual prompts greatly affects the final performance, i.e. at which layer we insert the prompts. For instance, the performance of BERT ranges from 64.97% to 69.54% on VQA2.0 with different placements. To address these shortcomings, we propose a novel multi- modal transfer learning approach for PLMs called Dynamic Visual Prompting (DVP) in this paper. Instead of using the entire image features, DVP first applies a cross-attention module to collect task-related and text-relevant visual infor- mation, which is further linearly projected onto the semantic space of PLMs as the prompt token. Meanwhile, to auto- matically set the optimal placement of these tokens, we also equip DVP with a novel search algorithm based on $k$ -armed bandit theory, which regards the token insertion as a policy action and effectively estimates the search weights via nu- merous single-shot trials (Zhou et al. 2020). In addition, to achieve parameter-efficient adaption of PLMs, we also com- bine DVP with the recently popular adapter approach (Sung, Cho, and Bansal 2022) for downstream VL tasks. Such a combination can not only greatly save the parameter expen- diture, but also enable PLMs to make a quick shift between NLP and VL tasks, as shown in Fig. 1. To validate DVP, we apply it to two representative PLMs, which are the encoder-based BERT (Devlinetal.2018) and the encoder-decoder based T5 (Raffeletal.2020). Extensive experiments are performed on a set of VL reasoning bench- marks including VQA2.0 (Goyal et al. 2017), GQA (Hud- sonandManning2019) and SNLI-VE (Xieetal.2019). To validate the generalization, we also apply DVP to a recent LLM called LLaMA (Touvron et al. 2023) on ScienceQA (Lu et al. 2022). The experimental results show that com- pared with the common visual prompting, DVP can save up to 80% computation while improving the performance sig- nificantly, e.g. +1.21% on VQA2.0 for BERT and +2.19% on GQA for T5. When combined with Adapter (Sung,Cho, and Bansal 2022), DVP can well maintain the overall per- formance of PLMs on these benchmarks, and it only needs to update about 5.0% -6.0% parameters of the model for VL adaptions. In addition to pre-training costs, DVP also ex- hibits much better efficiency than VLP models, e.g. 4.0G FLOPs of BERT-DVP v.s. 956.4G FLOPs of OSCAR (Li etal.2020). Meanwhile, when combined with Adapter, the updated parameters are only about 5.1% of OSCAR. These results well confirm our motivation about DVP. Overall, the contributions of this paper are three-fold: - We propose to directly adapt PLMs as a stand-alone model to VL tasks via inserting visual prompt tokens, which can avoid the building of heavy fusion networks and make use of the context reasoning ability of PLMs. - We propose a novel transfer learning method called D ynamic V isual P rompting (DVP) for efficient PLM adaption, which includes a cross-attention module to ob- tain compact visual tokens and a $k$ -armed bandit based search algorithm for automatic prompt placement. - DVP can reduce the computation of common visual prompting methods by up to 80% while achieving bet- ter performance on multiple benchmarks. Compared with VLP models, DVP can also help PLMs obtain competi- tive performance on VL tasks while saving the parame- ters and training overhead substantially. ## Related work With the great success of pre-training and fine-tuning paradigm in NLP, large-scale vision-language (VL) pre- training has also become a standard step in VL research (Chenetal.2023). In the early stage, the pre-trained vision encoders are introduced to provide the region knowledge. VisualBERT (Li et al. 2019), ViLBERT (Lu et al. 2019), LXMERT (TanandBansal2019), Uniter (Chenetal.2020), Oscar (Lietal.2020) use the pre-trained Faster-RCNN (Ren etal.2015) to extract region features and combine them with text features for deep fusion. To break through the limita- tion of object detectors in inference efficiency, PixelBERT (Huang et al. 2020), Grid-VLP (Yan et al. 2021), SOHO (Huangetal.2021) uses ResNet (Heetal.2016) to extract the grid features of the image. These methods not only real- ize end-to-end training of VLP models, but also achieve bet- ter performance on VL tasks. Although VLP models achieve excellent performance on downstream VL tasks, the PLMs are only used as a language encoder and still require another large fusion branch and expensive VL pretraining. Recently, Prompt tuning (Guetal.2021;Liuetal.2021; Wei et al. 2021; Han et al. 2022) aims to insert a set of prompt tokens into the input sequence of PLMs, thereby alleviating the difference between the data distributions of pre-training and downstream tasks (Liu et al. 2023b). To avoid laborious manual tuning, recent advances resort to learnable tokens for downstream tasks, which is also termed soft prompting (Lester, Al-Rfou, and Constant 2021). The great success of prompt tuning in NLP also sparks its ap- plication to computer vision (CV) and vision-language (VL) studies. In terms of VL studies, CoOp (Zhou et al. 2022b) keeps the parameters of CLIP unchanged, and puts learn- able parameters as soft prompting for the input text, which fully uses the zero-shot retrieval ability of CLIP. To solve the problem that CoOp is easy to overfit on the basic classes, CoCoOp (Zhou et al. 2022a) implements instance-adaptive
230600409/6
Table 3: Comparison between KAB-APP and the Manual search. “ Com-Prompt. ” refers to the common visual prompt- ing mentioned above. “ At 1 $st$ -layer ” denote that deploying DVP at the input layer of PLMs. \begin{tabular}{@{}cccccc@{}} \rightarrowprule \textbf{Model} & \textbf{Setting} & \textbf{FLOPs} & \textbf{\begin{tabular}[c]{@{}c@{}}VQAv2\end{tabular}} & \textbf{\begin{tabular}[c]{@{}c@{}}GQA\end{tabular}} & \textbf{\begin{tabular}[c]{@{}c@{}}SNLI-VE\end{tabular}} \midrule \multirow{4}{*}{BERT} & Com-Prompt. & 167.6G & 68.00 & 51.34 & \textbf{76.26} & At 1$st$-layer & 34.6G & 64.97 & 50.01 & 75.78 & Manual & 34.1G & \textbf{69.54} & 52.55 & 76.00 & KAB-APP & 34.0G & 69.21 & \textbf{52.55} & 74.76 \midrule \multirow{4}{*}{T5} & Com-Prompt. & 192.4G & 69.24 & 50.85 & \textbf{76.19} & At 1$st$-layer \tablefootnote{The first layer of the decoder of T5.} & 37.2G & 69.42 & 52.27 & 75.60 & Manual & 36.8G & 69.82 & 53.04 & 75.60 & KAB-APP & 36.8G & \textbf{69.82} & \textbf{53.04} & 75.60 \bottomrule \end{tabular} all image features is a reliable way of adapting PLMs to VL tasks, of which performance is consistent across tasks and models. However, compared with other solutions, its compu- tation is much more expensive, i.e. +4-+5 times. For [CLS] prompting, using a static visual prompt is inferior for most VL tasks, especially the ones that require fine-grained rea- soning, i.e. VQA2.0 and GQA. For the DVP methods, the dynamic prompt tokens can obtain better performance than the static [CLS] prompting in most cases, confirming our as- sumption about dynamic prompting. However, the increase of dynamic tokens does not always lead to better perfor- mance under all settings. Meanwhile, the effectivenesses of DVP on BERT and T5 are also different. Compared to T5, we directly place DVP at the input layer of BERT, which has yet to learn sufficient text semantics. This finding im- plies that the placement of prompt tokens is critical to DVP. Overall, these results confirm the feasibility of DVP, but it still needs great improvements in token placement. The impact of prompt placement. We further examine the impact of prompt placement via manual setting in Tab. 2. From this table, we can first confirm that the prompt place- ment is vital to performance. In terms of BERT, the per- formance of different DVP placements varies vastly, which changes are 4.57%, 2.54%, and 2.06% in VQA2.0, GQA, and SNLI-VE, respectively. In terms of T5, the performance change still exists but is less obvious, which are 0.6%, 1.18%, and 0.79%, respectively. To explain, we only try the placement of T5 from the decoding layers, where the textual semantic is well-built via its encoder branch. In this case, the quality of DVP is better ensured, and the similar cases can be also found in Tab. 1. Furthermore, we can see that the effect is also different on VL tasks. For instance, VQA performs better when the prompts are inserted to the higher layers and SNLI-VE does the opposite. For DVP, we specu- late that VQA firstly focuses on understanding the question itself, and then use its semantics to extract key information in the image to generate an effective prompt. In contrast, SNLI-VE may need more understanding of image seman- tics, and deploy it at a lower layer can better facilitate VL comprehension. Thus, we can conclude that the best inter- action of DVP varies for different tasks, which substantially confirms the necessity of automatic prompt placement. Effectiveness of KAB-APP. To validate KAB-APP, we compare it with the manual search in Tab. 3, which is ob- tained by training all possible insertions. It can be seen that the search results of KAB-APP is very close to that of man- ual search, which can be regarded as the upper-bound of DVP. On BERT, KAB-APP can achieve competitive perfor- mance against the manual search, although the best insertion layers are slightly different on VQA2.0, i.e. 11- $th$ and 10- $th$ layers, respectively. In terms of T5, of which text semantics are well learned, the best placements of KAB-APP and man- ual search are consistent on all tasks, greatly showing the effectiveness of our search algorithm. From Tab. 3, we can also see that via finding the optimal placement, the perfor- mance of DVP can be greatly improved. For instance, KAB- APP is better than using all visual prompts (Com-Prompt.) on VQA2.0 and GQA, while reducing the computation by 80.7%. Meanwhile, compared to the static placement, i.e. “ at 1- $st$ layer ”, the performance gains become more obvi- ous. These results well confirm our assumption and validate the effectiveness of KAB-APP. The combination with Adapter. To further improve ef- ficiency, we introduce Adapter (Sung, Cho, and Bansal 2022) to parameter-efficiently adapt PLMs into the down- stream tasks. As shown in the bottom two rows of Tab. 4, the adapter-based methods achieve similar performance to fine-tuning. Specifically, the performance on BERT is slightly improved compared to the fine-tuning mode, e.g. +1.07%, +0.92% and +0.94% on VQA2.0, GQA and SNLI- VE, respectively. In terms of T5, the DVP with adapters can achieve 99.3%, 99.8% and 98.8% performance of fine- tuning manner while only updating 5% parameters. From the above experiments, we can conclude that the proposed DVP method can easily cooperate with other PETL methods for adapting PLMs in a more efficient way. Comparison with visual-language pre-trained models. We also compare DVP with a bunch of VLP models on VL benchmarks in Tab. 4, which are often built with an- other deep fusion networks. Here, DVP is deployed accord- ing to the results of KAB-APP. Notably, compared to VLP models, DVP directly transfers PLMs into VL tasks with- out additional requirement of large-scale VL pre-training. When adapting pre-trained model into downstream tasks, DVP takes fewer computing resources especially when be- ing combined with the adapters, i.e. updates only 5.9% and 5.0% parameters of BERT and T5, respectively. When training similar parameters, the proposed DVP on BERT achieves competitive performance against ViLT on VQA2.0, GQA and SNLI-VE, respectively, while only taking 60.8% FLOPs. Compare to the SOTA method METER (Douetal. 2022), the proposed BERT-DVP $_{adp}$ method takes 2.2% trainable parameters and 14.1% FLOPs to achieve 70.28% performance on VQA2.0. Overall, these results confirm the feasibility of DVP in adapting PLMs to VL tasks. The generalization ability of DVP. In Tab. 5, we exam- ine the generalization of DVP to the recently proposed LLM called LLaMA (Touvronetal.2023) on ScienceQA. When applying DVP to LLaMA-7B, another VL pre-training is also not required, and updated parameters are only 9M, which is about 0.13% of LLaMA-7B. Compared with BLIP- 2 (Li et al. 2023) using Vicuna-7B as the base language
230600409/4
Figure2: Illustration of the proposed method. The right sub-figure illustrates the process of dynamic visual prompting (DVP), which can produce compact yet highly related visual tokens for prompting. The left sub-figure depicts the process of the proposed k-armed bandit based automatic prompt placement algorithm (KAB-APP). With a short search period, KAB-APP can find out the optimal deployment of DVP on different PLMs for different VL tasks. ## Automatic Prompt Placement In our experimental trials, we find that the model perfor- mance is disparate with the different placements of DVP. Meanwhile, for VL tasks with various definitions, a static setting is suboptimal to PLM adaption. To this end, we pro- pose a $k$ - A rmed B andit based A utomatic P rompt P lacement algorithm (KAB-APP) as shown in the left half of Fig. 2. Problem Definition. The target of KAB-APP is to auto- matically find out the insertion layer that can lead to the best performance of DVP. Thus, we consider the prompt place- ment as a policy action and model the search process as a $k$ -armed bandit problem. Concretely, we equally initialize the action preference for each layer of PLMs as $h \in \mathbb{R}$, that is, the probability of each layer being selected. During each search step, the policy for the $K$ - $th$ layer $\pi_t(K)$ is obtained by $$ \begin{aligned} \pi_t(K) &= \frac{e^{\mathbf{H}_t(K)}}{\sum_{m=1}^{M}e^{\mathbf{H}_t(l)}}, \end{aligned} $$ where $\mathbf{H}_t=[h_t^1,...,h_t^K,...,h_t^M]$ denotes the preference of each layer at $t$ -th setp. Based on Eq. 9, the action weight is then updated by $$ \begin{aligned} \Delta R &= R_t-R_b, \\ \mathbf{H}_{t+1}(K) &= \mathbf{H}_{t}(K)+\alpha\Delta R\pi_t(K)(1-\pi_t(K)), \end{aligned} $$ where $\alpha$ is the learning rate for updating $\mathbf{H}$, $R_t$ is the reward of inserting DVP to $K$ - $th$ layer at the $t$ - $th$ step, $R_b$ is the baseline reward obtained by $n$ sampling times. Here we use the average value of $n$ samplings’ rewards at the $t$ - $th$ step. Search Algorithm. The procedure of KAB-APP is de- scribed in Algorithm 1. Firstly, we establish an action space consisted of $M$ insertion layers for PLMs. Meanwhile, we equip all the possible insertion layers with independent cross-modal attention modules. At each training step, an inserted layer is randomly sam- pled from $M$, and this layer $K$ and the corresponding cross- attention module $\phi_K$ are activated for training. In a short training interval, we sample $n$ candidate insertion layers for validation by the RL policy $\pi_t$, and use the validation accu- racy or the power of the negative loss of $e$, as the rewards. Then we update preferences $\mathbf{H}$ by using the gradient update rules in Eq. 10. KAB-APP aims to search the best insertion layer for DVP on the language model. With the above search algorithm, the action weights can be fairly estimated via numerous single- shot validations (Zhou et al. 2020). Since the search space is limited, the whole process can be accomplished within a few periods. In practice, we control this search process to the limited steps, which only requires an additional 30-40% of the time of fine-tuning a PLM. ## Combination with Adapter We further combine DVP with an advanced parameter- efficient transfer learning approach called Adapter (Sung, Cho,andBansal2022;Luoetal.2023) to reduce the scale of training parameters. It can also help PLMs to make a quick shift between NLP and VL tasks. In practice, given a PLM, we can search the optimal DVP insertion layer $K$ with KAB- APP, and then use cross-modal attention layer $\phi$ to generate lightweight DVP at $K$ -th layer. Furthermore, we equipped language model with Adapters. The process above not only satisfies high parameter efficiency, but also attains optimal performance. ## Experiments ## Datasets VQA2.0 (Goyal et al. 2017) is a benchmark dataset that builds on VQA1.0 (Antol et al. 2015) and contains open- ended questions about images from MSCOCO (Ren,Kiros, andZemel2015). Following previous works (Yuetal.2019), we consider VQA2.0 as a classification task with 3129 an- swer categories. Following ViLT (Kim,Son,andKim2021), we use VQA2.0 train set and val set for fine-tuning while re- taining 1000 examples of the val set for validation.
230600409/1
# Adapting Pre-trained Language Models to Vision-Language Tasks via Dynamic Visual Prompting Shubin Huang 1, Qiong Wu 1, Yiyi Zhou 12 *, Weijie Chen 3, Rongsheng Zhang 3, Xiaoshuai Sun 12, Pre-trained language models (PLMs) have played an increas- ing role in vision-language (VL) learning, but they usually re- quire a deep multi-modal branch for VL reasoning, resulting in excessive computation and memory overhead. Recently, visual prompting is a feasible way to adapt PLMs to VL tasks, but we notice that the use of all visual tokens will greatly ex- ecrate the already high computation, and the token placement is also vital to performance. Based on these observations, we propose a novel transfer learning approach for PLMs in this paper, termed D ynamic V isual P rompting (DVP). Concretely, DVP first deploys a cross-attention module to obtain text- related and compact visual prompt tokens, thereby greatly reducing the input length of PLMs. To obtain the optimal placement, we also equip DVP with a reinforcement-learning based search algorithm, which can automatically merge DVP with PLMs for different VL tasks via a very short search pro- cess. In addition, we also combine DVP with the recently popular adapter approach to keep the most parameters of PLMs intact during adaption, which also help PLMs achieve a quick shift between single- and multi-modal tasks. We ap- ply DVP to two representative PLMs, namely BERT and T5, and a recent large language model called LLaMA. Extensive experiments are conducted on a set of VL reasoning bench- marks including VQA2.0, GQA, SNLI-VE and ScienceQA. The experimental results not only show the merits of DVP in performance and efficiency, e.g. +2.28% accuracy and -80% FLOPs on VQA2.0, but also confirm its superiority in adapt- ing pre-trained language models to VL tasks. Our code is anonymously released at https://github.com/hsb1357173526/ Dynamic Visual Prompting. ## Introduction Recent years have witnessed the rapid development of pre- trained language models (PLMs) (Devlinetal.2018;Lewis 2020), which have become the de facto standard in natu- ral language processing (NLP). Recently, the emergence of ChatGPT (OpenAI2022) and LLaMA (Touvronetal.2023) further confirms the significance of PLMs in exploring gen- eral artificial intelligence. The advent of PLMs also leads Figure 1: The application of the proposed D ynamic V isual P rompting (DVP) to pre-trained language models (PLMs). With low parameter and training expenditure, DVP can adapt PLMs to vision-language tasks and help them make a quick shift between single- and multi-modal tasks. to the prevalence of large-scale pre-training in the vision- language (VL) field (Lu et al. 2019; Tan and Bansal 2019; Lietal.2020;Shenetal.2021;Douetal.2022). However, directly adapting PLMs to VL tasks is pro- hibitively expensive. Above all, PLMs often act as a build- ing block for advanced VL models. Similar to vision mod- els (Ren et al. 2015; He et al. 2016; Dosovitskiy et al. 2020), PLMs are only used as a modal-specific encoder. For VL reasoning tasks like Visual Question Answering (VQA) (Goyal et al. 2017), the VL models still need to employ a deep fusion branch upon the multi-modal encoders, making the model extremely cumbersome. In addition to parameter redundancy, this deployment also undermines the ability of PLM in context reasoning since it only serves to embed the text words. To this end, the VL model often requires another large-scale pre-training on massive VL data (Luetal.2019; Son,andKim2021;Douetal.2022). In this case, we aim to directly adapt PLMs to vision- language tasks with low parameter and training overhead. A natural solution is to directly feed all the extracted visual features into PLMs as the prompt tokens. Here, we term it
230600409/3
by adding the vision feature of each image to learnable prompts, which makes prompt tuning more robust. The above methods mainly use text words or learnable tokens as the prompts for NLP and VL models. There are also some very recent works applying image information for VL task prompting. With an encoder-decoder architecture, SimVLM (Wang et al. 2021) uses the patch features extracted from images by ViT as prefix. Frozen (Tsimpoukellietal.2021) based on GPT only trains image encoder, and uses extracted visual features as visual prompt tokens for language model, which achieves excellent performance. Overall, existing vi- sual prompting methods often use the entire image features for VL adapting, while the critical issues of redundant visual information and excessive computation are still overlooked. ## Method ## Preliminary Before introducing our approach, we first recap the basic vi- sual prompting for PLMs. Concretely, given a PLM as $G(\cdot)$, and the image-text example as $(I,T)$, the target of adapting PLMs to VL tasks is to minimize the loss defined by $$ \setlength{\abovedisplayskip}{3pt} \setlength{\belowdisplayskip}{3pt} \begin{aligned} \operatorname*{argmin}_{\theta} \mathcal{L}\big(G(I,T|\theta)\big), \end{aligned} $$ where $\theta$ is the parameters of $G$. $\mathcal{L}$ is the objective function of the downstream task. For visual prompting, a natural so- lution is to use the all visual features as the prompt tokens, and then project them onto the semantic space of PLMs. However, in this solution, the length of input sequence $L$ is largely increased by the number of visual tokens $N$, which leads to the additional computations by $O(N^2+2NL)$. Be- sides, using all visual information for VL reasoning is often redundant to the model (Rao et al. 2021; Ryoo et al. 2021; Bolyaetal.2022). To alleviate the computational burdens of PLMs and re- duce the redundancy in visual prompts, we propose to use a lightweight network $\phi$ to generate compact visual prompts, thereby reducing the length of input sequence. The optimiza- tion objective can be further defined as $$ \setlength{\abovedisplayskip}{3pt} \setlength{\belowdisplayskip}{3pt} \begin{aligned} \operatorname*{argmin}_{\theta,\phi} \mathcal{L}\big(G(\phi(I,T),T|\theta)\big), \end{aligned} $$ which can obtain substantial improvements in both model inference and computation overhead. ## Dynamic Visual Prompting In this paper, we propose a Dynamic Visual Prompting (DVP) approach towards efficient VL adaption. In principle, DVP adopts cross-modal attention to dynamically generate visual prompts relevant to the input text and image, as shown in the right half of Fig. 2. Given an image $I$, we first use a frozen visual encoder to extract its features, denoted as $\mathbf{F}_v \in \mathbb{R}^{N \times d}$, where $N$ is the number of visual features and $d$ is the feature dimension. To obtain the compact visual tokens related to the input text, we use the text features in PLMs as the query vectors, $\mathbf{F}_t \in \mathbb{R}^{L \times d}$, where $L$ refers to the length of text sentences. With the projection weight matrices, $\mathbf{W}_Q \in \mathbb{R}^{d \times d}$, $\mathbf{W}_K \in \mathbb{R}^{d \times d}$, and $\mathbf{W}_V \in \mathbb{R}^{d \times d}$, for $Q, K, V$ transformations in our cross- modal attention module $\phi$. DVP is obtained by $$ \setlength\abovedisplayskip{6pt} \begin{aligned} \mathbf{F}_{DVP} = Softmax(\frac{\mathbf{F}_t\mathbf{W}_Q(\mathbf{F}_v\mathbf{W}_K)^T}{\sqrt{d}})\mathbf{F}_v\mathbf{W}_V. \end{aligned} $$ Here, we also employ multi-head attention mechanism, so the above formula can be further rewritten as $$ \begin{aligned} \mathbf{F}_{DVP} &= Concat(head_1, ..., head_h)\mathbf{W}_O, \\ head_i &= Softmax(\frac{\mathbf{F}_t\mathbf{W}^i_Q(\mathbf{F}_v\mathbf{W}^i_K)^T}{\sqrt{\frac{d}{n}}})\mathbf{F}_v\mathbf{W}^i_V, \end{aligned} $$ where $n$ is the number of attention heads, $\mathbf{W}_O \in \mathbb{R}^{d \times d}$ and $\mathbf{W}^i_Q, \mathbf{W}^i_K, \mathbf{W}^i_V \in \mathbb{R}^{d \times \frac{d}{n}}$ are the weight matrices. To this end, the length of generated $\mathbf{F}_{DVP}$ becomes $L$, and the computational complexity is reduced from $O\big((N+L)^2\big)$ to $O\big((L+L)^2\big)$ in calculating attention matrix, where $N$ is often larger than $L$ in VL tasks (Kim, Son, and Kim 2021; Shenetal.2021;Douetal.2022). However, the computation overhead is still expensive and we further reduce the length of prompt tokens to improve the efficiency by using the global features of PLMs as query vectors, such as [CLS] token of BERT, pooling feature of T5’s encoder or [EOS] token of LLaMA. For encoder-based language models like BERT, we define its layers as $\Theta=[\Theta_1,...,\Theta_{K},...,\Theta_{M}]$, where $K$ is the in- dex of insertion layer and $M$ refers to the number of layers. Here, $\mathbf{F}_t$ denotes the hidden features obtained before the $K$ - $th$ layer we want to insert, and we use its [CLS] token fea- tures $\mathbf{F}_t^{[CLS]}$ as query vector. Afterwards, DVP of BERT at insertion layer $K$ is obtained by $$ \setlength{\abovedisplayskip}{1pt} \setlength{\belowdisplayskip}{1pt} \begin{aligned} \mathbf{F}_t &= \Theta_{K-1}\big(\Theta_{K-2}...\Theta_1\big(E(T)\big)\big), \\ \mathbf{F}_{DVP} &= \phi(\emph{query}=\mathbf{F}_t^{[CLS]}, \emph{key}=\mathbf{F}_v, \emph{value}=\mathbf{F}_v), \end{aligned} $$ where $E$ denotes the text embedding. Next, $\mathbf{F}_{DVP}$ is con- catenated with $\mathbf{F}_t$ and sent to subsequent layers $$ [\mathbf{F}_{DVP};\mathbf{F}_t] = \Theta_M\big(\Theta_{M-1}...\Theta_{K}\big([\mathbf{F}_{DVP};\mathbf{F}_t]\big)\big) , $$ Finally, we use $\mathbf{F}_t^{[CLS]}$ to connect the specific VL task clas- sification head for prediction. For encoder-decoder based PLMs like T5, we define its encoder as $\Psi$, decoder layers as $\Theta=[\Theta_1,...,\Theta_{K},...,\Theta_{M}]$ and decoder input vector as $\mathbf{f} \in \mathbb{R}^{d}$. Here, $\mathbf{F}_t$ means the output of $\Psi$ and we utilize the pooling features of $\rho(\mathbf{F}_t)$ as text query vector, where $\rho$ means the operation of average mean pooling. Then $\mathbf{F_{DVP}}$ of T5 at insertion decoder layer $K$ is obtained by $$ \begin{aligned} \mathbf{F}_t &= \Psi\big(E(T)\big), \\ \mathbf{F}_{DVP} &= \phi(\emph{query}=\rho(\mathbf{F}_t), \emph{key}=\mathbf{F}_v, \emph{value}=\mathbf{F}_v). \end{aligned} $$ Afterwards, $\Theta_K$ ’s output is spliced with $\mathbf{F}_{DVP}$ and sent to the subsequent decoder layers: $$ \setlength{\abovedisplayskip}{1pt} \setlength{\belowdisplayskip}{1pt} \begin{aligned} \mathbf{f} &= \Theta_{K-1}\big(\Theta_{K-2}...\Theta_1\big(E(f)\big)\big), \\ [\mathbf{F}_{DVP};\mathbf{f}] &= \Theta_M\big(\Theta_{M-1}...\Theta_{K}\big([\mathbf{F}_{DVP};\mathbf{f}]\big)\big), \end{aligned} $$ Lastly, we use $\mathbf{f}$ attached to the classification head for pre- diction.
221016850/2
Third, based on the MIMIC-III dataset 3, we demonstrate the RAC model’s effectiveness in the most challenging full codes prediction testing set from inpatient clinical notes. The RAC model wins over all the previously reported SOTA results considerably. RAC establishes a new SOTA, considerably outperforming the current best Macro-F1 by 18.7%, and reaches past the human-level coding baseline. For further information on prediction results, see Table 1 and 2 of [1], as well as the Papers with Code’s leaderboard in the medical code prediction [5]. xRAC Framework and Human-Grounded Explainability Evaluation: Next, we present a general e x plainable R ead, A ttend, and C ode (xRAC) framework that generates evidentiary text snippets for a predicted code, which are oriented towards the needs of a deployment scenario. The first attention score-based approach, xRAC-ATTN, utilizes the label-wise attention mechanism first introduced in [4] to select key sentences for prediction decisions. In the second model-agnostic knowledge- distillation-based method, xRAC-KD, a large"teacher" RAC model is distilled into a collection of"student" linear models in post-hoc manner without sacrificing much accuracy of the teacher model while retaining many advantages of linear models including explainability and smaller model size which is beneficial for deployment. More methodological details are available in Section 2 of [2]. Then, a simplified but thorough human-grounded evaluation with two groups of annotators, one group (Group A) with and one group without (Group B) medical coding expertise 4, is conducted to evaluate the explainability of our xRAC framework, xRAC-ATTN and xRAC-KD. Both groups are given the same question sheet, which has explanation text snippets extracted to support the appearance of the predicted codes. Each annotator must select one of the three choices that are highly informative, informative, and irrelevant. More information on the evaluation task design and execution can be found in Section 3.2 of [2]. We find that the proposed xRAC framework can be easily explainable and the supporting evidence text highlighted by xRAC-ATTN is of higher quality than xRAC-KD even though xRAC-KD has potential advantages in production deployment scenarios. One can see from Table 2 in [2] that there is much larger gap in xRAC-KD between Group A and Group B than between xRAC-ATTN. This implies that xRAC-ATTN is a more viable choice than xRAC-KD to extract a text snippet from clinical notes to support code prediction. However, Table 3 in [2] shows that the consistency score measured by Jaccard Similarity between two groups is lower than 40% even with xRAC-ATTN. This suggests that the automated extraction system must still rely on professional coders’ feedback, and there is room remaining to improve for a lay person without expertise to correctly code. We believe that this is a very meaningful step toward the goal of developing accurate, explainable and automated ML systems for medical code prediction from clinical notes. To our knowledge, most previous works are still in early stages in terms of providing textual references and explanations of the predicted codes and no studies to date have thoroughly analyzed in this regard, especially for complex transformer-based models such as the RAC model. ## Conclusion In this study, we present for the first time a human-coding baseline for medical code prediction on the subsampled MIMIC-III-full-label inpatient clinical notes testing set task. We have developed an attention-based RAC model that sets the new SOTA records, and the resulting RAC model outperforms the human-coding baseline to a great extent on the same task. The performance improvements can be attributed to effectively learning the common embedding space between the clinical note and medical codes by utilizing attention mechanisms that efficiently address the severe long-tail sparsity issues. Additionally, we present a xRAC framework to obtain supporting evidence text from clinical notes that justify predicted medical codes from medical code prediction systems. We have demonstrated that the proposed xRAC framework may help even complex transformer-based models to attain high accuracy with a decent level of explainability (which is of high value for deployment scenarios) through qualitative human-grounded evaluations. We also show for the first time that, given the current state of explainability methodologies, using the proposed explainable yet accurate medical codes prediction system still requires professional coders’ expertise and competencies. 3 The MIMIC-III Dataset (MIMIC v1.4) is a freely accessible medical database containing de-identified medical data of over 40,000 patients staying in the BIDMC between 2001 and 2012. 4 Group A had two annotators without medical coding experience and Group B had six certified professional coders.
221016850/1
# Medical Codes Prediction from Clinical Notes: From Human Coders to Machines Byung-Hak Kim AKASA, Inc. South San Francisco, CA [email protected] ## Introduction Prediction of medical codes from clinical notes is a practical and essential need for every healthcare delivery organization within current medical systems 1. Automating annotation will save significant time and excessive effort that human coders spend today. However, the biggest challenge is directly identifying appropriate medical codes from several thousands of high-dimensional codes from unstructured free-text clinical notes. This complex medical codes prediction problem from clinical notes has received substantial interest in the NLP community, and several recent studies [1,6,3,8] have shown the state-of-the-art (SOTA) code prediction results of full-fledged deep learning-based methods. This progress raises the fundamental question of how far automated machine learning (ML) systems are from human coders’ working performance, as well as the important question of how well current explainability methods apply to advanced neural network models such as transformers. This is to predict correct codes and present references in clinical notes that support code prediction, as this level of explainability and accuracy of the prediction outcomes is critical to gaining trust from professional medical coders. ## Explainable and Accurate Medical Codes Prediction RAC Model and Human-Level Coding Baselines: We first present our R ead, A ttend, and C ode (RAC) model to learn the assignment mappings of medical codes that can process unstructured clinical notes and attend to text areas annotating medical codes. The main building blocks of RAC are built upon by connecting convolved embeddings with Transformer-encoder based self-attention and code-title guided attention modules. The architecture details can be found in Section 4 of [1]. In principle, we look at the medical codes prediction problem as a set-to-set assignment learning problem from the set of input sentence vectors to the set of code labels and employ the problem’s unique permutation equivariant property 2 in the design. Second, the human coding baseline is estimated via a primitive web interface built using the Label Studio [7]. Given the reference diagnosis and procedure codes that Beth Israel Deaconess Medical Center (BIDMC) assigned, we have two professional CPC certified coders independently code the different set of notes and evaluate where their total annotations have differed from the references. The intention is to provide the same possible conditions as for the RAC system. We found that estimated inter-coder agreement rates (i.e., human coding baseline) were not as high as we initially thought and are far exceeded by the RAC system’s 3.9 times higher rate in Micro-Jaccard similarity. Section 3 of [1] conatins the evaluation design details and baseline results. 1 A human coder or health care provider scans the medical documentation in electronic health records, identifying essential information and annotating codes for that particular treatment or service. With a wide range of medical services and providers (primary care clinics, specialty clinics, emergency departments, mother-baby units, outpatient and inpatient units, etc.), the complexity of human coders’ tasks increases as the medical industry advances, while productivity standards decrease as charts take more time to review. 2 Consider permuting the notes’ sentence vectors; medical code outputs will be the same but permuted.
190800995/5
In Fig. 2, we show the numerically obtained value of $\overline{C}_\infty M/4$ for various $k$ and $q$. The alternating pattern ex- pected from Eq. (13) is clearly visible [panels (a) to (c)]. While the numerical value of the nonzero plateau differs from expectation when $q \ll k$ (and $k-q \ll k$, not shown), Eq. (13) gives an accurate prediction when $\binom{k}{q}/\binom{k}{\lfloor k/2 \rfloor}$ is close to one (with $\lfloor \ldots \rfloor$ the floor function). To investi- gate this further, in panels (d) and (e), we show $C_\infty M/4$ versus $k$. The growth with $k$ is slower for $q=5$ than for $q=3$, which is in turn slower than the almost linearly growing $q=1$ case [46]. To summarize, we have shown that supersymmetry is (almost) always present in the SYK model with generic four-body interactions. It is only absent in those classes without particle-hole symmetry, i.e., in classes AI and AII. The type of SUSY, in particular the number $\mathcal{N}_\text{loc}$ of local supercharges and their symmetry properties fol- low a pattern that finds a natural interpretation when $\Gamma_{j\le \mathcal{N}_\mathrm{loc}}$ are viewed as emergent Majorana fermions in a one-dimensional topological phase with SYK model boundary physics. These SUSY features all link directly to features in time-dependent correlation functions of fermion-parity-odd observables. For $q$ -body retarded cor- relation functions, this includes the shape of the ramp in the short-time regime, due to a link between $\mathcal{N}_\text{loc}$ and the Dyson index $\beta$; and the value of the long-time plateau due to the imprint of how $\Gamma_j$ transforms under $T_\pm$ on its mi- croscopic Majorana structure. These $q$ -body correlation functions, even with large $q$, can be measured in digital quantum simulation of the SYK model [53]. The single- particle Green’s function ($q=1$) is accessible through scanning tunneling spectroscopy [20, 21]. We stress that the features in the correlation functions are dynamical consequences of SUSY, which are less frequently consid- ered than ground-state consequences [54, 55]. We thank David Tong for helpful discussions. This work was supported by the ERC Starting Grant No. 678795 TopInSy. ## Appendix A: Eightfold symmetry classification of the SYK model In the main text, we make extensive use of the eight- fold symmetry classification of the SYK model with four-body interactions. This classification, introduced in Ref. 46, reveals an Altland-Zirnbauer structure behind the period-eight pattern of Wigner-Dyson classes high- lighted in Ref. 22. The considerations of both Refs. 22 and 46 are based on the work of Fidkowski and Kitaev on the topological classification of interacting fermions in one dimension [45], which Ref. 22 noted to apply to the SYK model upon viewing it as existing at the end of a one-dimensional fermionic topological phase. Here we summarize the key ideas in Ref. 46 and highlight certain aspects relevant for the main body of the paper. The eightfold symmetry classification is based on the presence of antiunitary symmetries and their relation to fermion parity. First note that the SYK Hamilto- nian [Eq. (2) in the main text] conserves fermion parity, $[H,P]=0$ with $$ P = \begin{cases} i^{k/2} \gamma_1 \gamma_2 \cdots \gamma_k & \text{even $k$} \\ i^{(k+1)/2} \gamma_1 \gamma_2 \cdots \gamma_k \gamma_\infty & \text{odd $k$} \end{cases} $$ where the Majorana $\gamma_\infty$ ensures a well-defined fermion parity for odd $k$, but does not contribute to local opera- tors, including the Hamiltonian. All Majorana operators including $\gamma_\infty$ are odd in fermion parity and thus satisfy $\{P,\gamma_q \} =0$. When $k$ is odd, the Hamiltonian also com- mutes with the operator $$ Z = - i^{(k-1)/2} \gamma_1 \gamma_2 \cdots \gamma_k $$ that in turn commutes with all local Majorana opera- tors $[Z,\gamma_{q \neq \infty}]=0$, but anticommutes with the additional Majorana at infinity $\{Z,\gamma_{\infty} \}=0$ and accordingly with fermion parity $\{Z, P\} = 0$. Depending on the number of Majorana operators, it may be possible to find local unitary operators $C_\pm$ that satisfy [56] $$ C_\pm \gamma_{q \neq \infty}^* C_\pm^\dagger = \pm \gamma_{q \neq \infty} . $$ In other words, the antiunitary operator $T= C_+ \mathcal{K}$ commutes with all local Majorana operators, whereas $A = C_- \mathcal{K}$ anticommutes with them. For even $k$, it is always possible to find both operators $C_+$ and $C_-$ [56]. Using Eq. (A1) we conclude $$ T P T^{-1} = (-1)^{k/2} P, $$ i.e., $T$ commutes with fermion parity when $k\mod 4 = 0$ and anticommutes with fermion parity when $k\mod 4 = 2$. Thus, when both operators commute, $T$ is an antiu- nitary operator that maps each parity subblock of the Hamiltonian to itself, otherwise, the parity subblocks are exchanged by $T$. We refer to the first case as time- reversal symmetry: The level spacing statistics of each subblock are determined by the presence of $T$ and by the sign $T^2 = \pm 1$. We refer to the second case as particle-hole symmetry, as the parity sectors are exchanged. Particle- hole symmetry sets correlations across different parity sectors, a feature we exploit in the main text. In the fol- lowing, we distinguish the two cases by the notation $T_+$ for time-reversal ($[T_+,P]=0$) and $T_-$ for particle-hole symmetry ($\{T_-,P \} =0$), the same notation we use in the main text. For odd $k$, only one of two operators $C_\pm$ is local. In particular, when $k =4n+1$, only $C_+$ is local, and when $k = 4n+3$, only $C_-$ is local (with integer $n$). When only $C_-$ is local, we can use fermion parity to define $T= P C_- \mathcal{K}$, an antiunitary operator that commutes with
190800995/7
Ensemble-averaged $q$ -body correlation function in symmetry classes AI [$k \mod 8 =0$, panels (a) and (b)] and AII [$k \mod 8 =4$, panels (c) and (d)]. The color code de- notes the different values of $k$ and $q$; cf. the inset in panel (a). Here, $q=5$ serves is a representative example for $q=4n+1$, and $q=3$ for $q=4n+3$. The correlation function decays to zero in all cases shown here. Only when $k=12$ [panels (c) and (d)] do the two parity sectors show correlations, which is a small-size effect. All results are averaged over a large ensemble, ranging from $768$ (for $k=20$, $q=5$) to $2^{17}$ (for $k=8$) realizations of the couplings $J_{qrst}$. Error bars are ei- ther smaller than the line width (for small $k$) or smaller than the disorder-induced fluctuations of the lines (for large $k$). For completeness, we give these operators in terms of the eigenstates. To give explicit relations, we need to fix certain phase relations between the eigenstates. Defining time-reversal symmetry as $$ T_+ \ket{\psi_{2\mu+1}^+} = \ket{\psi_{2\mu}^+} $$ fixes the phase relation between Kramers doublets that we denote by $\ket{\psi_{2\mu}^p}$ and $\ket{\psi_{2\mu+1}^p}$. This choice sets $\varepsilon_{2\mu} = \varepsilon_{2\mu+1}$. We fix the phase between opposite parities by choosing $$ Z \ket{\psi_{\mu}^p} = \ket{\psi_{\mu}^{-p}} . $$ In this basis, the supercharges defined above read $$ \begin{aligned} \Gamma_1 &= \sum_{p\mu} \left(\ket{\psi_{2\mu}^p} \bra{\psi_{2\mu+1}^{-p}} + \ket{\psi_{2\mu+1}^p} \bra{\psi_{2\mu}^{-p}} \right) \\ \Gamma_2 &= -i \sum_{p\mu} \left(\ket{\psi_{2\mu}^p} \bra{\psi_{2\mu+1}^{-p}} - \ket{\psi_{2\mu+1}^p} \bra{\psi_{2\mu}^{-p}} \right) \\ \Gamma_3 &= \sum_{p\mu} \left(\ket{\psi_{2\mu}^p} \bra{\psi_{2\mu}^{-p}} - \ket{\psi_{2\mu+1}^p} \bra{\psi_{2\mu+1}^{-p}} \right) . \end{aligned} $$ Their product is accordingly $$ \begin{aligned} \Gamma_4 &= -i \Gamma_1 \Gamma_2 \Gamma_3 = \sum_{p\mu} \ket{\psi_{\mu}^p} \bra{\psi_{\mu}^{-p}} = Z \end{aligned} $$ Ensemble-averaged $q$ -body correlation function in symmetry classes BDI ($k\mod 8=1$) and CII ($k\mod 8 = 5$). In panels (a) and (b), we show different system sizes in class BDI, where the different colors denote $k$ and $q$; cf. inset in panel (d). As in Fig. 3, $q=5$ serves is a representative example for $q=4n+1$, and $q=3$ for $q=4n+3$. In panels (c) and (d), we show different systems sizes in class CII. The dotted lines show the expectation for $\overline{C}_{q,\infty}$ based on random matrix theory; cf. Eq. (13) in the main text. All results are averaged over a large ensemble, ranging from $192$ (for $k=21$, $q=5$) to $2^{14}$ (for $k=9$) realizations of the couplings $J_{qrst}$. Similarly to Fig. 3, error bars are either smaller than the line width (for small $k$) or smaller than the disorder-induced fluctuations of the lines (for large $k$). and the nonlocal supercharge $$ \begin{aligned} \Gamma_\infty &= -i \sum_\mu \left(\ket{\psi_{\mu}^+} \bra{\psi_{\mu}^{-}} - \ket{\psi_{\mu}^-} \bra{\psi_{\mu}^{+}} \right) . \end{aligned} $$ Writing the fermion parity $P$ in terms of the eigenstates $$ \begin{aligned} P &= \sum_\mu \left(\ket{\psi_{\mu}^+} \bra{\psi_{\mu}^{+}} - \ket{\psi_{\mu}^-} \bra{\psi_{\mu}^{-}} \right) \end{aligned} $$ and using $P=-i Z \gamma_\infty$ confirms that $\Gamma_\infty \equiv \gamma_\infty$. ## Appendix C: Correlation functions a. Symmetry classes AI and AII. Classes AI ($k \mod 8 = 0$) and AII ($k \mod 8 =4$) are the two classes without SUSY. In these classes, parity sectors are un- correlated, hence, upon evaluating Eq. (9) of the main text, we expect to see neither a ramp structure nor a plateau at long times $t \gg 1/\Delta_\infty$. To support this expec- tation with numerical evidence, we show the ensemble- averaged $q$ -body correlation function at infinite tempera- ture [Eq. (10) in the main text] in Fig. 3. In Fig. 3 (a), we show the $(q=5)$ -body correlation function for $k=8,16$
190800995/2
\begin{tabular}{c|cccccccc} \rightarrowprule $k \mod 8$ & 0 & 1 & 2 & 3 & 4 & 5 & 6 & 7 \colrule Label & AI & BDI & D & DIII & AII & CII & C & CI \colrule $T_+^2$ & $+1$ & $+1$ & $0$ & $-1$ & $-1$ & $-1$ & $0$ & $+1$ $T_-^2$ & $0$ & $+1$ & $+1$ & $+1$ & $0$ & $-1$ & $-1$ & $-1$ \botrule \end{tabular} TABLEI. Time-reversal symmetry $T_+$ and particle-hole sym- metry $T_-$ in the SYK model. The symmetries may be absent (denoted by $0$), or present and square to $-1$ or $+1$. Wigner-Dyson way highlighted in Ref. 22. In fact, as we now briefly review, it gives rise to the eight real Altland-Zirnbauer classes. For even $k$, either time- reversal symmetry $T_+$ or particle-hole symmetry $T_-$ is present. For odd $k$, both $T_+$ and $T_-$ are present; in this case $T_+ \gamma_{\infty} T_+^{-1} = (-1)^{(k+1)/2}\gamma_{\infty}$. Their product, the unitary operator $Z= T_+T_-$, equals the product of all local Majorana operators up to a complex phase and corresponds to a chiral symmetry [45, 46]. A key feature of $Z$, which we will use repeatedly for diagnosing local- ity, is $[Z,\gamma_{q\neq\infty}]=0$ and $\{Z,\gamma_\infty\}=0$. The squares $T_\pm^2$ vary with $k$ and label the eight real Altland-Zirnbauer classes [48]. While the Altland-Zirnbauer and Wigner- Dyson picture give the same level spacing statistics, the former also takes cross-parity correlations into account. We summarize the symmetry classification in Table I and review it in detail in Appendix A. SUSY is known to imply a degeneracy between the par- ity sectors [36]: the supercharges $Q_a$ exchange bosonic states with parity eigenvalues $p=+1$ and fermionic states with parity eigenvalue $p=-1$ [36]. Thus, the su- percharges anticommute with fermion parity, $\{P, Q_a \} =0$. The presence of particle-hole symmetry also guar- antees degeneracy between parity sectors, which as we now note, also implies SUSY. Parity degeneracy directly follows from particle-hole symmetry because $\ket{\psi^p_\mu}$ and $T_- \ket{\psi^p_\mu}$ have the same energy $\varepsilon_\mu=\varepsilon_\mu^p= \varepsilon_\mu^{-p}$ (since $[T_-,H]=0$), but opposite parity ($\lbrace T_-,P \rbrace = 0$). Therefore, $\ket{\psi_\mu^p} \bra{\psi_\mu^{-p}}$ is an odd-parity zero mode, i.e., an operator that commutes with the Hamiltonian, but anticommutes with fermion parity [46]. This in turn implies SUSY: The operator $\tilde{Q}_\mu = \sqrt{\varepsilon_\mu} \ket{\psi_\mu^+} \bra{\psi_\mu^{-}}$ sat- isfies $\tilde{Q}_\mu \tilde{Q}_\mu^\dagger = \varepsilon_\mu \ket{\psi_\mu^+} \bra{\psi_\mu^+}$ and $\tilde{Q}_\mu^\dagger \tilde{Q}_\mu = \varepsilon_\mu \ket{\psi_\mu^{-}} \bra{\psi_\mu^{-}}$, and hence the linear combinations $Q_{1,\mu} = \tilde{Q}_\mu + \tilde{Q}_\mu^\dagger$ and $Q_{2,\mu} = i (\tilde{Q}_\mu - \tilde{Q}_\mu^\dagger)$ are Hermitian, anticommute, and square to $\varepsilon_\mu$ times the projector on the two parity- degenerate states. Consequently, the two supercharges $$ \begin{aligned} Q_1 = \sum_\mu (\tilde{Q}_\mu + \tilde{Q}_\mu^\dagger) , & & Q_2 = -i \sum_\mu (\tilde{Q}_\mu - \tilde{Q}_\mu^\dagger) \end{aligned} $$ satisfy Eq. (1) and anticommute with $P$. Particle-hole symmetry is present unless $k=4n$. Thus, all but two of the symmetry classes are supersymmetric. Given the presence of six supersymmetric classes, there are a number of questions regarding the interplay of SUSY and the symmetry classification. How does $\mathcal{N}$ depend on the symmetry class? How do $Q_j$ transform under $T_\pm$ and how does this translate to the structure of the supercharges? We next turn to these questions. We start with counting $\mathcal{N}$. A direct approach is based on counting level degeneracies. This follows from the ob- servation that the “ spectrally flattened ” Hermitian su- percharges $\Gamma_{j}=Q_{j}/\sqrt{H}$ satisfy $$ \begin{aligned} \{\Gamma_{j},\Gamma_{k}\} = 2\delta_{jk}, & & [H,\Gamma_{k}] = 0 , & & \{P,\Gamma_{k}\}=0. \end{aligned} $$ They are thus many-body zero mode forms of Majorana fermions. An even $\mathcal{N}$ of such zero modes give rise to a $2^{\mathcal{N}/2}$ -dimensional fermionic degeneracy space for each of the $\varepsilon_\mu$ with one of the $|\psi_\mu^p\rangle$ chosen as “ vacuum ”. (With a suitable choice, the $\Gamma_j$ -fermion parity of an eigenstate matches the state’s physical fermion parity.) This pro- cedure is similar in spirit to the standard construction of supermultiplets [49]. For the six supersymmetric SYK classes, a twofold degeneracy is guaranteed by $T_-$ and a further twofold (Kramers) degeneracy is present when- ever $T_+^2=-1$, resulting in an overall fourfold degeneracy. This suggests $\mathcal{N}=2$, except for DIII and CII where this count gives $\mathcal{N}=4$. What this counting does not address is how many $\Gamma_j$ (and hence $Q_j$) are local to the SYK model. Next we investigate this to obtain the decompo- sition $\mathcal{N}=\mathcal{N}_\text{loc}+\mathcal{N}_\infty$ with $\mathcal{N}_\text{loc}$ counting the number of supercharges involving only $\gamma_{q\neq\infty}$. We first discuss the symmetry classes D and C before demonstrating the im- plications of locality in classes BDI and CI. For brevity, we derive the supercharges in classes DIII and CII with $T_-^2 = -1$ in Appendix B and only summarize the results here. We begin with classes D and C. Here $k$ is even hence all $\gamma_q$ are local. Therefore, our argument above applies directly: we find $\mathcal{N}=\mathcal{N}_\text{loc}=2$. All the other supersym- metric classes have odd $k$, thus potentially $\mathcal{N}\neq\mathcal{N}_\text{loc}$ due to $\gamma_\infty$. As we shall see, in all of these classes $\mathcal{N}=\mathcal{N}_\text{loc}+1$ with $\mathcal{N}$ following its degeneracy-based value above. This is intuitive because $\gamma_\infty\equiv \Gamma_\infty$ auto- matically satisfies Eq. (5) (in particular, it anticommutes with any local parity-odd operator), thus $\mathcal{N}_\text{loc}$ is at most $\mathcal{N}-1$. To formally establish $\mathcal{N}_\text{loc}$, and the transforma- tion of $\Gamma_{j}$ under $T_\pm$, we work in the energy eigenbasis, $H = \diag (\lbrace \varepsilon_\mu \rbrace) \otimes \openone_{2^{\mathcal{N}/2}}$, with $P = \openone_{M/2}\otimes\tau_3$. (Here and below, $\tau_j$ and $\sigma_j$ are Pauli matrices; $\tau_j$ act in parity grading and $\sigma_j$ in the space of Kramers doublets, where applicable. We will often omit trivial tensor factors.) In this basis, class D (C) has particle-hole symmetry [up to a phase $\diag (\lbrace \exp(i\varphi_\mu) \rbrace)$ omitted here and be- low] $T_- = \tau_{1(2)} \mathcal{K}$ (with $\mathcal{K}$ for complex conjugation); this follows from $T_-^2=\pm1$ and parity being the only degener- acy, $\openone_{2^{\mathcal{N}/2}}=\tau_0$. We have $\Gamma_{1,2}=\tau_{1,2}$, which correspond to the two supercharges introduced in Eq. (4) [50]. To study classes BDI and CI, we focus on a degeneracy space with energy $\varepsilon_\mu$ and first establish the form of $T_\pm$ and thus $Z$ in this space. $T_+^2=+1$ implies that parity is again the only degeneracy, so $T_- = \tau_{1(2)} \mathcal{K}$ in class BDI (CI). $T_+\ket{\psi_\mu^p} \propto \ket{\psi_\mu^p}$ implies that the most general form is $T_+=\exp(i\varphi_{\mu}\tau_{3})\mathcal{K}$. With a suitable choice of
190800995/6
all fermions; when $C_+$ is local, we can use $T=C_+ \mathcal{K}$. This implies that $T$ and $P$ commute $$ \begin{aligned} T P T^{-1} &= (-1)^{(k+1)/2} \gamma_1 \gamma_2 \cdots \gamma_k T \gamma_\infty T^{-1} \\ &= P, \end{aligned} $$ where we chose $T \gamma_\infty T^{-1} = (-1)^{(k+1)/2} \gamma_\infty$ for conve- nience. We thus identify $T = T_+$. Due to the presence of $Z$, we can define a second antiunitary operator $T_-$ via $T_- = T_+^{-1} Z$ such that $Z = T_+ T_-$ as quoted in the main text. Since $Z$ anticommutes with fermion parity, $T_-$ also anticommutes with it, $\{T_- ,P \} =0$. The squares of $T_+$ and $T_-$ can be determined from their explicit construction in terms of Majorana operators [45]. Without loss of generality, we choose the Majorana oper- ators such that $\gamma_{2n+1}^* = \gamma_{2n+1}$ are real while $\gamma_{2n}^* = -\gamma_{2n}$ are purely imaginary. When $k$ is even, the basis choice introduced above re- sults in $$ C_+ = \begin{cases} \gamma_2 \gamma_4 \cdots \gamma_{k} & ~ k = 4 n \\ \gamma_1 \gamma_3 \cdots \gamma_{k-1} & ~ k = 4 n +2 , \end{cases} $$ thus, $C_+$ is always the product of $k/2$ Majorana oper- ators. In both cases, $C_+ = C_+^*$ is real ($C_+$ is either the product of only real Majorana operators or the product of an even number of purely imaginary operators). When $k=4n$, we have $$ \begin{aligned} T_+^2 = C_+^ 2 = \gamma_2\gamma_4 \cdots \gamma_k \gamma_2 \gamma_4 \cdots \gamma_k , \end{aligned} $$ giving $T_+^2 = (-1)^{k/4}$, and when $k=4n+2$, $$ \begin{aligned} T_+^2 = C_+^ 2 = \gamma_1\gamma_3 \cdots \gamma_{k-1} \gamma_1 \gamma_3 \cdots \gamma_{k-1} , \end{aligned} $$ which gives $T_+^2 = (-1)^{(k-2)/4}$. When $k$ is odd, we need to distinguish the two cases $k=4n+1$ and $k=4n+3$. For $k=4n+1$ we find $$ C_+ = \gamma_2 \gamma_4 \cdots \gamma_{k-1} . $$ Since $k$ is odd, the nonlocal Majorana $\gamma_\infty^* = -\gamma_\infty$ is purely imaginary, giving $$ C_+ \gamma_\infty^* C_+^\dagger = -(-1)^{(k-1)/2} \gamma_\infty = (-1)^{(k+1)/2} \gamma_\infty $$ where we used that $C_+$ contains $(k-1)/2$ Majorana operators $\gamma_{q \neq \infty}$. Accordingly, we identify $T_+ = C_+ \mathcal{K}$ and conclude $T_+^2 = (-1)^{(k-1)/4}$. We further identify $T_- = T_+^{-1} Z$ with $T_-^2 = T_+^2 T_+ Z T_+^{-1} Z$. Using $T_+ Z T^{-1} = (-1)^{(k-1)/2} Z = Z$ and $Z^2 = 1$, we realize that $T_-^2 = T_+^2$. For $k = 4n+3$ we find $$ C_- = \gamma_1 \gamma_3 \cdots \gamma_{k} , $$ which is a product of $(k+1)/2$ Majorana operators. Since $C_- \mathcal{K}$ and $P$ both anticommute with all local Ma- jorana operators, the antiunitary operator $P C_- \mathcal{K}$ com- mutes with them. Again, $\gamma_\infty^* = - \gamma_\infty$, which gives $$ P C_- \mathcal{K} \gamma_\infty (P C_- \mathcal{K})^{-1} = (-1)^{(k+1)/2} \gamma_\infty, $$ where we used that $C_-$ is the product of $(k+1)/2$ local Majorana operators and $\{P,\gamma_\infty \} = 0$. Thus, we iden- tify $T_+ = P C_- \mathcal{K}$. The square $T_+^2 = P C_- P^* C_- = (-1)^{k+1} C_-^2$ and $C_-^2 = (-1)^{(k+1)/4}$, such that $T_+^2 = (-1)^{(k+1)/4}$. Accordingly, $T_-^2 = T_+^2 T_+ Z T_+^{-1} Z = (-1)^{(k-1)/2} T_+^2 = - T_+^2$. We summarize these results in Table I in the main text. ## Appendix B: Supercharges in classes DIII and CII In this section, we construct the supercharges in classes DIII and CII along the lines of the same strategy that we followed in the main text for the other classes. We show and discuss how the supercharge structure summarized in the main text arises in these classes. In classes DIII and CII, time reversal symmetry with $T_+^2 = -1$ implies Kramers degeneracy: The states $\ket{\psi_\mu^p}$ and $T_+ \ket{\psi_\mu^p}$ are orthogonal. Together with parity degen- eracy this gives a fourfold degeneracy of energy eigenval- ues and $\mathcal{N}=4$. In its diagonalized form, the Hamiltonian reads $H=\diag (\{\varepsilon_\mu \}) \otimes \openone_{2^{\mathcal{N}/2}}$ with $\openone_{2^{\mathcal{N}/2}}=\tau_0\sigma_0$, where the matrices $\tau_\mu$ act in parity grading space and $\sigma_\nu$ in the space of Kramers doublets (hence $\tau_\mu \sigma_\nu$ is a shorthand for the Kronecker product $\tau_\mu \otimes \sigma_\nu$; additional tensor prod- ucts with the identity are implied). Without loss of generality, we can choose $Z=\tau_1$. The four $\Gamma_j$ satisfying Eq. (5) in the main text are now the familiar Dirac matrices, chosen as $\Gamma_j=\tau_1\sigma_j$ ($j=1,2,3$) and $\Gamma_\infty=\tau_2$. Checking the (anti)commutation with $Z$ we indeed find that only $\Gamma_{1,2,3}$ are local. Conversely, identifying $\Gamma_\infty\equiv \gamma_\infty$ and requiring that it transforms under $T_+$ as $T_+ \gamma_\infty T_+^{-1} = (-1)^{(k+1)/2} \gamma_\infty$ sets $$ T_+= \begin{cases} i\tau_3\sigma_2\mathcal{K} & \text{class DIII} \\ i\sigma_2\mathcal{K} & \text{CII} \end{cases} $$ in this basis. The form of particle-hole symmetry $T_-$ follows from $T_+ T_-= Z$. The product $\Gamma_4 = -i \Gamma_1 \Gamma_2 \Gamma_3$ is also Hermitian, par- ity odd, local, and linearly independent of $\Gamma_{1,2,3}$. Fur- thermore, $Q_4= \diag (\lbrace \sqrt{\varepsilon_\mu} \rbrace) \otimes \Gamma_4$ squares to the Hamil- tonian. It is, however, not a supercharge because $\Gamma_4$ does not anticommute with $\Gamma_{1,2,3}$. (Nevertheless, $\Gamma_4$ con- tributes to correlation functions, as discussed in the main text.) We thus find $\mathcal{N}_\text{loc}=3$. Using the explicit form of $T_+$ obtained above and $T_- = T_+^{-1} Z$, we find $$ T_\pm \Gamma_{j \le \mathcal{N}_\mathrm{loc}} T_\pm^{-1} = \begin{cases} +\Gamma_{j \le \mathcal{N}_\mathrm{loc}} & \text{class DIII} \\ -\Gamma_{j \le \mathcal{N}_\mathrm{loc}} & \text{class CII.} \end{cases} $$ This implies that $$ T_\pm \Gamma_4 T_\pm^{-1} = - \Gamma_4 = \begin{cases} - \Gamma_4 & \text{class DIII} \\ + \Gamma_4 & \text{class CII.} \end{cases} $$
190800995/4
$q$ -body time-dependent correlation function at infinite temperature, averaged over an ensemble of up to $2^{16}$ Gaussian distributed $J_{qrst}$, for classes (a) D, (b) C, (c) DIII, and (d) CI. The different colors denote different $k$ and $q$, cf. inset in (a), the dashed lines represent $q=3$ and the solid lines $q=5$. The ramp shape follows the Dyson index $\beta$ and hence links to the number of supercharges. The long-time plateau $\overline{C}_{q,\infty}$ is studied in Fig. 2. Error bars are either smaller than the line width (for small $k$) or smaller than the disorder-induced fluctuations of the lines (for large $k$). time-dependent $q$ -body correlation function $$ C_{q}^+ (t) = - i \Theta (t) \frac{1}{\binom{k}{q}} \sum_{a,n_a =q} \langle \lbrace \Upsilon_a (t) , \Upsilon_a (0) \rbrace \rangle , $$ where $\langle \ldots \rangle$ denotes thermal average. Although the sig- natures we reveal are present at any temperature, we find an especially transparent relationship at infinite temper- ature, where the correlation function reads $$ \begin{aligned} C_{q}^+ (t) =& -i \Theta (t) \frac{1}{\binom{k}{q}} \sum_{a,n_a =q} \frac{1}{M} \sum_{p\mu\nu} \left| \bra{\psi_\mu^p} \Upsilon_a \ket{\psi_\nu^{-p}} \right|^2 \\ & \times 2 \cos \left(t \left(\varepsilon_\mu^p - \varepsilon_\nu^{-p} \right) \right). \end{aligned} $$ When $t \gg 1/\Delta_\infty$, terms with $\varepsilon_\mu^p \neq \varepsilon_\nu^{-p}$ give a quickly os- cillating contribution $\delta C_q^+ (t)$ that averages to zero. Only states with $\varepsilon_\mu^p = \varepsilon_\nu^{-p}$ give a time-independent contribu- tion $C_{q,\infty}$. Thus, $C_q^+ (t) = -i \Theta(t) [C_{q,\infty} + \delta C_q^+ (t)]$ with $$ \!\!\!\! C_{q,\infty}=\frac{1}{\binom{k}{q}}\sum_{a,n_{a}=q}\frac{2}{M}\sum_{\mu} \tr \Upsilon_{a\mu}^{2},\quad\Upsilon_{a\mu}=P_{\mu}\Upsilon_{a}P_{\mu}, $$ where in converting the equal-energy sum to a trace, we introduced the projection $P_{\mu}$ to the eigenspace with energy $\varepsilon_{\mu}$ and used that $\Upsilon_{a}$ is Hermitian and parity odd. Next we convert Eq. (11) into a sum over $\Gamma_{j<\infty}$. We start by expanding $\Upsilon_{a\mu}=\sum_{j<\infty}y_{j}P_{\mu}\Gamma_{j}$ (with real $y_j$) Normalized plateau $\overline{C}_{q,\infty} M/4$ of the $q$ -body correla- tion function, averaged over an ensemble of up to $2^{14}$ Gaussian distributed $J_{qrst}$. The color encodes the number $k$ of Majo- ranas, cf. panels (d) and (e). In all classes, $\overline{C}_\infty M/4$ alternates with $q$ approximately as predicted in Eq. (13); the agreement is excellent when $\binom{k}{q}/\binom{k}{\lfloor k/2 \rfloor}$ is close to one. In panel (d), we show that $\overline{C}_\infty M/ (4c)$ [with $c$ the random matrix expectation based on Eq. (13)] increases as a function of $k$, but with a rate that decreases upon increasing $q$ [panel (e)]. Statistical error bars are smaller than the marker size. which holds as within an eigenspace, the (projected) operators $P_{\mu}\Gamma_{j<\infty}$ form a basis for local, Hermitian, parity-odd operators. If $\Upsilon_{a}$ transforms the same (op- posite) way to $\Gamma_{j}$ under $T_{\pm}$ then generically $y_{j}\neq0$ ($y_{j}=0$). Now using the trace-orthogonality of the $\Gamma_{j<\infty}$ and $\tr \Gamma_{j}^{2} = 2^{\mathcal{N}/2}=\mathcal{N}$ (for $\mathcal{N}=2,4$), we find $$ C_{q,\infty}=\frac{1}{\binom{k}{q}} \sum_{a,n_{a}=q} \frac{2}{M\mathcal{N}} \sum_{\mu} \sum_{j<\infty} \left[\tr(P_{\mu}\Upsilon_{a}P_{\mu}\Gamma_{j}) \right]^{2}. $$ A simple estimate for $C_{q,\infty}$ can be given assuming that expanding $P_\mu\Gamma_{j<\infty}=\sum_{a} v_{\mu j,a}\Upsilon_{a}$ results in random co- efficients $v_{\mu j,a}$ subject only to normalization and symme- try constraints. Denoting such a random vector average by $\overline{(\ldots)}$, we find $$ \frac{\overline{C}_{q,\infty} M}{4}=\begin{cases} \frac{\mathcal{N}}{\beta}\mathcal{N}_{\text{loc}} & q:\,\Upsilon_{a}\triangleq\Gamma_{j\leq\mathcal{N}_{\text{loc}}},\\ \frac{\mathcal{N}}{\beta}\delta_{\beta,4} & \text{otherwise}, \end{cases} $$ where $\Upsilon_{a} \triangleq \Gamma_j$ here means that $\Upsilon_{a}$ transforms the same way as $\Gamma_j$ under $T_\pm$. Thus, each $\Gamma_{j<\infty}$ give the same con- tribution to $\overline{C}_{q,\infty}$ when they contain $q$ -Majorana terms and zero otherwise. The nonzero value for $\beta=4$ when $\Upsilon_{a}\not\triangleq \Gamma_{j\leq \mathcal{N}_\text{loc}}$ arises due to $\Gamma_4$ since $\Gamma_4\not\triangleq\Gamma_{j\leq \mathcal{N}_\text{loc}}$. Con- sidering the Majorana structure of $\Gamma_j$ in Table II, Eq. (13) translates to an alternating pattern of $\overline{C}_{q,\infty}$ as $q$ is var- ied in a given symmetry class, with complementary $\overline{C}_{q,\infty}$ values for classes with opposite $s\mathcal{N}_\text{loc}$.
190800995/1
# Supersymmetry in the nonsupersymmetric Sachdev-Ye-Kitaev model Jan Behrends 1 and Benjamin B ´e ri 1,2 Supersymmetry is a powerful concept in quantum many-body physics. It helps to illuminate ground state properties of complex quantum systems and gives relations between correlation func- tions. In this work, we show that the Sachdev-Ye-Kitaev model, in its simplest form of Majorana fermions with random four-body interactions, is supersymmetric. In contrast to existing explicitly supersymmetric extensions of the model, the supersymmetry we find requires no relations between couplings. The type of supersymmetry and the structure of the supercharges are entirely set by the number of interacting Majorana modes, and are thus fundamentally linked to the model’s Altland- Zirnbauer classification. The supersymmetry we uncover has a natural interpretation in terms of a one-dimensional topological phase supporting Sachdev-Ye-Kitaev boundary physics, and has conse- quences away from the ground state, including in $q$ -body dynamical correlation functions. The Sachdev-Ye-Kitaev (SYK) model [1, 2] is a toy model that provides insight into diverse physical phe- nomena, ranging from the holographic principle [3 – 5] to quantum chaos [6 – 11], and non-Fermi liquid behavior of strongly correlated electron systems [12 – 18]. Similar to black holes, it is believed to scramble quantum informa- tion with maximal efficiency [17, 19]. The simplest variant of the SYK model describes $k$ Majorana fermions that interact through a random four- body term [2]. Its proposed physical realizations include mesoscopic systems based on Majorana fermions in vor- tices or quantum dots [20, 21], or the ends of a one- dimensional topological phase [22]. Various generalizations of the SYK model have been considered, including models with $n$ -body inter- actions [23, 24] and supersymmetric extensions [25 – 29]. Typically, exact supersymmetry (SUSY) requires fine- tuning of the parameters [30 – 33]. In the supersymmetric SYK extensions, this fine-tuning corresponds to requiring certain relations between different couplings [25]. In this work, we show that already the simplest four- body SYK model, without any fine-tuning, is supersym- metric for all but two values of $k \mod 8$. The type of SUSY depends only on $k$. The supercharges will be shown to relate to ramps and long-time plateaus in time- dependent correlation functions [34], which thus provide signatures of SUSY far from equilibrium. In particu- lar, we find that the number of supercharges is linked to the presence and nature of time-reversal symmetry and is reflected in the ramp shape [35]. We also show that the number and structure of supercharges set the plateau features in $q$ -body time-dependent correlation functions. Throughout this work, we focus on SUSY in the sense of supersymmetric quantum mechanics [30, 31, 36 – 42]. SUSY is characterized by $\mathcal{N}$, the number of mutually anticommuting Hermitian fermionic supercharges that square to the Hamiltonian [36] $$ \begin{aligned} \lbrace Q_a , Q_b \rbrace = 2 H \delta_{ab} , & & [H,Q_a] = 0 . \end{aligned} $$ The Hamiltonian we consider describes four-body in- teractions between $k$ Majorana modes [2] $$ H = \sum_{t =0}^{k-1} \sum_{s=0}^{t-1} \sum_{r=0}^{s-1} \sum_{q=0}^{r-1} J_{qrst} \gamma_q \gamma_r \gamma_s \gamma_t + E_0 $$ with real (as required by Hermiticity) but otherwise structureless couplings $J_{qrst}$, and the constant $E_0$ that ensures positive energies. The Hermitian Majorana op- erators $\gamma_q = \gamma_q^\dagger$ satisfy the anticommutation relation $\lbrace \gamma_q , \gamma_r \rbrace = 2 \delta_{qr}$, and span an $M$ -dimensional Hilbert space with $M=2^{\lceil k/2 \rceil}$. Since each term in the Hamiltonian (2) contains an even number of Majoranas, it conserves fermion parity $P$, given by $$ P = \begin{cases} i^{k/2} \gamma_1 \gamma_2 \ldots \gamma_k & \text{even $k$} \\ i^{(k+1)/2} \gamma_1 \gamma_2 \ldots \gamma_k \gamma_\infty & \text{odd $k$.} \end{cases} $$ To work in a Hilbert space with well-defined fermion parity, the additional Majorana $\gamma_\infty$ “ at infinity” must be included when $k$ is odd [45]. The operator $\gamma_\infty$ is not local to the SYK model; considering, e.g., a realization in a superconducting vortex [20], it represents a degree of freedom with support far away from the vortex where the SYK Majoranas $\gamma_{j\neq \infty}$ reside. Like the local Ma- joranas, $\gamma_\infty$ is Hermitian and satisfies $\{\gamma_q ,\gamma_r \} = 2 \delta_{qr}$. Since $[H,P]=0$, all eigenstates of $H$ can by labeled by their parity eigenvalue $p=\pm 1$, giving $H \ket{\psi_\mu^p} = \varepsilon_\mu^p \ket{\psi_\mu^p}$ and $P \ket{\psi_\mu^p} = p \ket{\psi_\mu^p}$. The number of interacting Majorana modes, specif- ically $k\mod 8$, sets the model’s antiunitary symme- tries [22, 45 – 47]. These come in two variants $T_\pm$, antiu- nitary operators satisfying $T_\pm \gamma_{q\neq \infty} T_\pm^{-1} = \gamma_{q\neq \infty}$. They further satisfy $T_\pm P T_\pm^{-1}= \pm P$. We refer to $T_+$ as time- reversal symmetry because it commutes with fermion parity and hence sets correlations within a parity sector. Conversely, we call $T_-$ particle-hole symmetry. Crucially for this work, since $T_-$ flips fermion parity, its presence implies correlations between parity sectors. The consideration of both $T_+$ and $T_-$ implies [46] a classification with more structure than the threefold
190800995/3
the relative phases between the two parity sectors we can thus use $Z=T_+ T_-=\tau_1$; in this basis $T_+=\mathcal{K}$ ($T_+=\tau_3\mathcal{K}$) for class BDI (CI). The two $\Gamma_j$ satisfying Eq. can again be chosen as $\Gamma_{1,2}=\tau_{1,2}$. However, checking the (anti)commutation with $Z$ we find that only $\Gamma_1$ is local. Conversely, we can identify $\Gamma_2\equiv\Gamma_\infty\equiv \gamma_\infty$; this is consistent both with $\gamma_\infty$ itself satisfying Eq. (5) and its transformation under $T_+$. We thus find $\mathcal{N}_\text{loc}=1$. In classes DIII and CII, we find $\mathcal{N}_\mathrm{loc} =3$ local super- charges as we show in detail in Appendix B. The spec- trally flattened supercharges can be written as Kronecker products $\Gamma_j =\tau_1 \sigma_j$. Their product $\Gamma_4 = -i \Gamma_1 \Gamma_2 \Gamma_3 = \tau_1$ is also local, but does not anticommute with $\Gamma_{j \le \mathcal{N}_\mathrm{loc}}$; it does, however, contribute to correlation function, as we discuss in the following. As in classes BDI and CI, the nonlocal supercharge is $\Gamma_\infty = \tau_2$. The values $\mathcal{N}_\text{loc}$, together with the sign $s$ in $T_\pm \Gamma_{j\leq\mathcal{N}_\text{loc}}T_\pm^{-1} = s\Gamma_{j\leq\mathcal{N}_\text{loc}}$ have a natural interpretation if one views the SYK model as arising at the end of a one-dimensional topological phase in class BDI [22, 45]. These systems admit a $\mathbb{Z}_8$ classification: At one of their ends, they have $k_s$ Majoranas satisfying $T_\pm \gamma_q T_\pm^{-1}=s\gamma_q$; the topological index is $\nu= (k_+-k_-) \mod 8$. Thus, the eight topological classes can be labeled by $\nu = 0,1,2,3,4,-3,-2,-1$ with the integers counting the number and sign of unpaired Majoranas. In the SUSY classes, we find the same pattern for $s\mathcal{N}_\text{loc}$ against $k \mod 8$ ($T_\pm \gamma_{q\neq \infty} T_\pm^{-1}=\gamma_{q \neq \infty}$ implies $k_+=k$, $k_-=0$), see Table II. The $\mathcal{N}_\text{loc}$ supercharges $\Gamma_{j\leq\mathcal{N}_\text{loc}}$ can thus be viewed as the many-body emergence of the minimal num- ber and type of unpaired Majoranas consistent with $k$. Next we turn to the structure of the supercharges in terms of the Majorana fermions $\gamma_q$. For this, we employ another operator basis of the Hilbert space, the products of $n_a$ Majorana operators $\gamma_{q\neq\infty}$ $$ \Upsilon_a = i^{n_a (n_a -1)/2} \gamma_{i_1 (a)} \gamma_{i_2 (a)} \cdots \gamma_{i_{n_a} (a)} $$ with $i_j (a) \neq i_{j'} (a)$. $\Upsilon_a$ are Hermitian, unitary, and or- thonormal with respect to the trace, $\tr \left[\Upsilon_a \Upsilon_b \right] /M = \delta_{ab}$. In total, there are $2^{k}$ local operators $\Upsilon_a$ [51]. As we aim to expand $\Gamma_{j\neq \infty}$, i.e., Hermitian odd-parity operators in terms of $\Upsilon_a$, we use only those $\Upsilon_a$ with odd $n_a$, and use only real expansion coefficients. Both time-reversal and particle-hole symmetry have the same (anti-) commutation properties when acting on $\Upsilon_a$. Since $T_\pm \gamma_q T_\pm^{-1} = \gamma_q$, only the phase of $\Upsilon_a$ [cf. Eq. (6)] may change when applying $T_\pm$, giving $$ T_\pm \Upsilon_a T_\pm^{-1} = (-1)^{n_a (n_a -1)/2} \Upsilon_a. $$ That is, $T_\pm$ and $\Upsilon_a$ commute when $n_a = 4n + 1$, and anti- commute when $n_a = 4n +3$. This, together with $v_{j,a}\in \mathbb{R}$ below, implies that, when expanding the supercharges, $$ \Gamma_j = \sum_a v_{j,a} \Upsilon_a ,\quad \sum_a v_{j,a}^2=1, $$ only terms with $n_a = 4 n+1$ contribute to $\Gamma_j$ when $[T_\pm ,\Gamma_j] = 0$, and only terms with $n_a = 4n+3$ contribute \begin{tabular}{c|cccccc} \rightarrowprule $k \mod 8$ & 1 & 2 & 3 & 5 & 6 & 7 \colrule Label & BDI & D & DIII & CII & C & CI $\beta$ & $1$ & $2$ & $4$ & $4$ & $2$ & $1$ \colrule $s\mathcal{N}_\text{loc}$ & $1$ & $2$ & $3$ & $-3$ & $-2$ & $-1$ $\Gamma_{j\leq\mathcal{N}_\text{loc}}$ & $4n+1$ & $4n+1$ & $4n+1$ & $4n+3$ & $4n+3$ & $4n+3$ \colrule $\Gamma_{4}$ & & & $4n+3$ & $4n+1$ & & \botrule \end{tabular} TABLE II. The Dyson index $\beta$, number $\mathcal{N}_\text{loc}$, signature $T_\pm \Gamma_{j\leq\mathcal{N}_\text{loc}}T_\pm^{-1}=s\Gamma_{j\leq\mathcal{N}_\text{loc}}$, and the Majorana fermion struc- ture of $\Gamma_{j\neq \infty}$. (The supercharges $Q_j$ have the same proper- ties, since T ± HT − 1 = H.) In the Majorana expansion of $\Gamma_j$, only those $\Upsilon_a$ with $n_a = 4n+1$ or $n_a = 4n+3$ contribute; the two options are shown in the last two rows of the table. The horizontal line visually distinguishes $\Gamma_4$ from the three supercharges because it does not anticommute with them. A blank entry indicates that $\Gamma_4$ does not exist in these classes. when $\lbrace T_\pm,\Gamma_j \rbrace =0$. In classes DIII and CII we also consider $\Gamma_4=-i\Gamma_1\Gamma_2\Gamma_3$ whose transformation proper- ties follow from those of $\Gamma_{1,2,3}$. The resulting expansion structure is summarized in Table. II. Having discussed the interplay of SUSY and the sym- metry classification, we now identify signatures of SUSY, $\mathcal{N}_\text{loc}$, and the supercharge structure in various observ- ables. A simple link between $\mathcal{N}_\text{loc}$ and observables exists due to the fact that the number of different $\Gamma_{j\leq \mathcal{N}_\text{loc}}$ and their linearly independent odd-parity products, i.e., in- cluding $\Gamma_4$ in classes CII and DIII, equals the degrees of freedom $\beta$ (i.e., the Dyson index linked to $T_+$ [52]) of the Hamiltonian’s off-diagonal matrix elements. In fact, the most general Hermitian linear combinations of these $\Gamma_j$ have the same type of offdiagonals, up to an imaginary unit, as the Hamiltonian: real for $\beta=1$ (classes BDI and CI), complex for $\beta=2$ (classes D and C) and real quaternion for $\beta=4$ (classes DIII and CII). In the SUSY classes, the value of $\beta$ sets the en- ergy level correlations, including the long-range spec- tral rigidity, across opposite parity sectors (these are uncorrelated without SUSY) which lead to “ ramps ” in time-dependent correlation functions of parity- odd ob- servables. (For single-Majorana examples see Refs. 34 and 46.) These ramps occur at time scales below $2\pi$ times the inverse mean level spacing $1/\Delta_\infty$, and have $\beta$ - dependent shape [35]. In particular, the ramp connects to a long-time plateau smoothly when $\beta=1$, sharply when $\beta=2$, and with a kink when $\beta=4$. In Fig. 1, we show ensemble-averaged $q$ -body correlation functions [Eq. (10) below] in classes D, C, DIII, and CI. For completeness, we show the correlation function in the remaining sym- metry classes, including those that do not support SUSY, in Appendix C. Besides this direct correspondence between the super- charges and ramp structure, we additionally find more subtle consequences of SUSY: The long-time ($t \gg 1/\Delta_\infty$) plateau in $q$ -body correlation functions is also related to the number and structure of the supercharges, cf. Fig. 2. To quantify this relationship, we consider the retarded
221207153/5
Gilles Chabrier etal.:ImpactofanewH / Heequationofstateontheevolutionofmassivebrowndwarfs \begin{tabular}{lcccccc} \hline\hline Object & & $M$ [${\mathrm{M_{Jup}}}$] & age [Gyr] & $\mathrm{T_{eff}}$ [K] & $\log(L/L_\odot)$ & $R/R_\odot$ \hline HD4113C & Observations$^{1,8}$ & 66$\pm 5$ & 5$^{+1.3}_{-1.7}$ & 500-600 & $-6.30\pm 0.22$ & % & B2003 & 63 & 5 & 1120 & -5.06 & 0.079 % & & 66 & 5 & 1520 & -4.50 & 0.081 & P2020 & 63 & 5 & 1108 & -5.06 & 0.080 & & 66 & & 1149 & -4.99 & 0.080 & {\bf present} & 63 & 5 & 1081 & -5.11 & 0.079 & & 66 & & 1144 & -5.02 & 0.079 % Eps Indi Ba & Observations$^2$ & 75$\pm 0.82$ & {\rm old} & 1320 & -4.70 & 0.083 Eps Indi Ba & Observations$^2$ & 66.92$\pm 0.36$ & 3.5$^{+0.8}_{-1.0}$ & & & & Observations$^5$ & 68$\pm 0.9$ & & 1312$\pm9$ & -4.70$\pm 0.02$ & & P2020 & 66-67 & 3.0 & 1324-1395 & -4.72 - -4.63 & 0.083 \ & & & 3.5 & 1262-1325 & -4.81 - -4.73 & 0.082 \ & & & 3.8 & 1233-1293 & -4.85 - -4.78 & 0.081 \ & {\bf present} & 66-67 & 2.8 & 1321-1354 & -4.74 - -4.70 & 0.081 \ & & & 3.0 & 1291-1321 & -4.78 - -4.74 & 0.081 \ & & & 3.5 & 1241-1265 & -4.86 - -4.83 & 0.080 \ Eps Indi Bb (or C) & Observations$^2$ & 53.25$\pm 0.29$ & 3.5$^{+0.8}_{-1.0}$ & & & & Observations$^{5}$ & 53$\pm 0.3$ & & 975$\pm 11$ & -5.23$\pm 0.02$ & & P2020 & 53 & 3.0 & 1071 & -5.08 & 0.084 \ & & & 3.5 & 1023 & -5.17 & 0.083 \ & & & 3.8 & 996 & -5.21 & 0.083 \ & {\bf present} & 53 & 2.8 & 1060 & -5.10 & 0.083 & & & 3.0 & 1042 & -5.13 & 0.083 & & & 3.5 & 998 & -5.22 & 0.083 \ & & & 3.8 & 973 & -5.26 & 0.082 GL 758 B & Observations$^3$ & 42$^{+19}_{-7}$ $(>30)$ & 1-6\,({\rm older?}) & 650 & -6.07$\pm 0.03$ & & Observations$^4$ & $38.1^{+1.7}_{-1.5}$ & $\gtrsim 6$ & & & & P2020 & 37 & 5 & 680 & -5.84 & 0.087 & & & 8 & 600 & -6.07 & 0.086 & {\bf present} & 37 & 5 & 673 & -5.86 & 0.087 & & & 8 & 593 & -6.09 & 0.085 WISE J0720-0846B & Observations$^6$ & 66$\pm 4$ & $>$ a few & 1250$\pm$40 & -4.82$\pm 0.07$ & % & B2003 {\bf 2x-CHECK} & 67 & 5 & 1270 & -4.83 & 0.078 & P2020 & 66 & 3 & 1324 & -4.72 & 0.082 & & & 4 & 1216 & -4.88 & 0.082 & {\bf present} & 66 & 3 & 1291 & -4.78 & 0.081 & & & 4 & 1201 & -4.92 & 0.080 2M0805+48 & Observations$^7$ & 66$^{+5}_{-14}$ & $\ge 4$ & & & 2M1059-21 & Observations$^7$ & $67^{+4}_{-5}$ & & & & & P2020 & 66 & 5 & 1149 & -4.99 & 0.080 & & & 10 & 964 & -5.33 & 0.077 & {\bf present} & 66 & 5 & 1142 & -5.02 & 0.079 & & & 10 & 975 & -5.33 & 0.076 Gliese 229 B & Observations$^8$ & 71.4$\pm 0.6$ & $<$10 & & -5.208$\pm 0.007$ & & P2020 & 70.5 & 8 & 1171 & -4.99 & 0.078 & & & 9 & 1140 & -5.04 & 0.077 \ & & & 10 & 1109 & -5.09 & 0.077 \ & {\bf present} & 70.5 & 8 & 1133 & -5.06 & 0.077 & & & 9 & 1103 & -5.11 & 0.076 \ & & & 10 & 1075 & -5.16 & 0.076 \ & & 71.5 & 10 & 1110 & -5.10 & 0.076 \hline \hline \end{tabular} Table2. E ff ective temperature, luminosity and radius of the various BDs examined in the text obtained with the P2020 and present models for the observed masses and ages. Observations: $^1$ Cheetham et al. (2018), $^2$ Chen et al. (2022), $^3$ Bowler et al. (2018), $^4$ Brandt et al. (2019), $^5$ Cardoso (2012), $^6$ Dupuy et al. (2019), $^7$ Sahlmann et al. (2020), $^8$ Brandt et al. (2021). Models are Phillips et al. (2020, P2020) and present calculations. We have taken 1 ${\mathrm{M_{Jup}}}$ = $9.5\times 10^{-4}\,{\mathrm{M}_\odot}$ in the evolutionary models. Using high-contrast imaging with the SPHERE instrument at the VLT, Cheetham et al. (2018) obtained the first images of the cold BD HD4113C (Sp = T9), which is part of a dynami- cal system with a M-dwarf companion. The dynamical mass is ${\mathrm{M_{dyn}}}=66\pm5\,{\mathrm{M_{Jup}}}$, while comparison of the observed spectrum of HD4113C with atmospheric models (Morley et al. 2012, 2014 and Tremblin et al. 2015) yields $\mathrm{T_{eff}} \sim500$ -600 K, $\log\, g$ = 4.5-5.0 and a radius $R\sim 1.4$ -1.5 ${\mathrm{R_{Jup}}}$, much larger than predicted for old, high-mass BDs. Using stellar evolution models (Mowlavi et al. 2012), the derived age of the parent star is found to be $\tau=5^{+1.3}_{-1.7}$
221207153/7
Gilles Chabrier etal.:ImpactofanewH / Heequationofstateontheevolutionofmassivebrowndwarfs Fig.4. Left: cooling curves for a $67\,{\mathrm{M_{Jup}}}$ (top) and a $53\,{\mathrm{M_{Jup}}}$ (bottom) brown dwarf, respectively, representative of the $\epsilon$ Ind Bab system. Solid line: present models, based on the CD21 EOS; dashed line: Phillips et al. (2020) models, based on the CMS19 EOS. Right: brown dwarf isochrones typical of the inferred age of the $\epsilon$ Ind Bab system, calculated with the present (red) and Phillips et al. (2020) (black) models. Fig.5. Left: e ff ective temperature as a function of mass for massive BDs for 3 isochrones calculated with the present models, based on the CD21 EOS and ATMO atmosphere models, the Phillips et al. (2020) models, based on the CMS19 EOS and ATMO atmosphere models and the Bara ff e et al. (2003) models, based on the SCvH EOS and the COND atmosphere m odels. Right: luminosity as a function of mass for 3 isochrones for BDs with dynamical mass measurements. References for the data are given in Table 2. All models have an equivalent helium mass fraction $Y_{eq} = 0.292$. al. (2019). The BD companion has a mass ${\mathrm{M_{dyn}}}=66\pm4\,{\mathrm{M_{Jup}}}$, an e ff ective temperature $\mathrm{T_{eff}}= 1250\pm40$ K and a luminosity $\log (L/L_\odot)=-4.8\pm0.15$. This suggests an age older than a few Gyr, consistent with the age estimates for the primary star. As shown in Table 2, both the P2020 and present models yield a nearly perfect agreement with the observational determinations for an age $\sim$ 3-4 Gyr, with the new ones predicting a slightly younger age (more rapid cooling). For the primary star, WISE J0720-0846A, as noted in Dupuy et al. (2019), models (BHAC 2015) overestimate the luminosity for its mass (${\mathrm{M_{dyn}}}=99\pm6\,{\mathrm{M_{Jup}}}$), or conversely underestimate its mass for its luminosity, at about 2 $\sigma$ (see their Table 5 and Fig. 7). As mentioned earlier, the new CD21 EOS does not significantly change this analysis for such high (stellar) masses (see Table 1). It is worth noting that the mass and age determinations of WISE J0720-0846B are very similar to the ones of the BD HD4747B ($66\pm 3\,{\mathrm{M_{Jup}}}$, $2.9^{+0.5}_{-0.4}$ Gyr, $\log (L/L_\odot)=-4.55\pm0.08$; Brandt et al. 2021, Table 10), for which the present models give a very good agreement (see Table 2). ## 3.5. 2M1059 and 2M0805 Sahlmann et al. (2020) have measured the complete astromet- ric orbits for the systems 2M0805 + 48 and 2M1059-21. They
221207153/2
Gilles Chabrier etal.:ImpactofanewH / Heequationofstateontheevolutionofmassivebrowndwarfs Fig.1. Interior $T$ - $\rho$ profiles in the hydrogen phase diagram for di ff erent astrophysical bodies at about 5 Gyr, as labelled in the figure (red lines) and for a 0.075 ${\mathrm{M}_\odot}$ object at 10 $^8$, $8\times$ 10 $^8$ and $8\times$ 10 $^{9}$ yr (blue dotted, short-dashed and long-dashed lines, re- spectively). The SCvH, CP and QMD labels stand for Saumon, Chabrier & vanHorn (1995) EOS, Chabrier & Potekhin (1998) EOS and QMD simulations (see Chabrier et al. (2019) and Chabrier & Debras (2021) for further details on the line sym- bols). eral perspective, it should be mentioned that the aforementioned recent analysis of Brandt et al. (2021) suggests another, di ff er- ent issue for young ($<$ 1 Gyr) and low-mass ($\lesssim 40\,{\mathrm{M_{Jup}}}$) BDs. In that case, the trend is the opposite: models underpredict lu- minosities for given mass and age, or equivalently overestimate (resp. underestimate) the mass (resp. the age) for the correct age (resp. mass). For objects in-between, mass $\sim$ 40-70 ${\mathrm{M_{Jup}}}$, age $\sim$ 1- 5 Gyr, models agree well ($<1\sigma$) with observations (see Table 10 of Brandt et al. (2021)). This points to two di ff erent issues in BD cooling theory in the extreme mass and age limits. In the present paper, we will focus on the first issue, for which we show that part of the solution lies in the EOS. In contrast, the second problem stems more likely from remaining uncertainties in BD atmosphere models. ## 2. Impact of the new equations of state on the internal structure, cooling history and hydrogen-burning limit Figure 1 displays temperature-density profiles of various astro- physical bodies, from 1 ${\mathrm{M}_\odot}$ to 1 ${\mathrm{M_{Jup}}}$ for an age of about 5 Gyr, in the phase diagram of hydrogen, for the $T$ - $\rho$ range covered by the CMS19 and Chabrier & Debras (2021, CD21) EOSs. This diagram is similar to the ones portrayed in these papers for H and He and we refer the reader to these papers for more details. The key issue here is that, as illustrated by the central part of the diagram labelled QMD (for Quantum Molecular Dynamics), all BD cooling tracks enter the domain where interactions be- tween H and He species can no longer be ignored, as assumed in the so-called ideal (or ’additive’) volume law approximation, used in CMS19 and Saumon et al. (1995, SCvH) (see Fig. 1 of CD21 for further details). Recently, Chabrier & Debras (2021) took into account the impact of these interactions on the thermo- dynamic properties of the H / He mixture by incorporating in the EOS the QMD calculations performed by Militzer & Hubbard (2013). The figure also displays the cooling sequence of a 0.075 ${\mathrm{M}_\odot}$ object at 10 $^8$, $\sim$ 10 $^9$ and $\sim$ 10 $^{10}$ yr, respectively. As seen in the figure (see also the Conclusion of CMS19), essentially all objects below $0.1\,{\mathrm{M}_\odot}$ older than about $\sim 0.1$ Gyr will enter the domain where H-He interactions cannot be ignored and will af- fect to some degree their mechanical and thermal properties, thus their structure and evolution. It is the aim of the present paper to examine in detail this impact. The left panel of Fig. 2 displays the evolution of the cen- tral density, $\rho_c$, temperature, $T_c$, and degeneracy parameter, $\psi_c=(T_c/T_F)\approx 3.314\times10^{-6}T_c(\mu_e/\rho)^{2/3}$,where $T_F$ and $\mu_e$ denote the electron Fermi temperature and mean molecular weight, respec- tively, of a 0.075 ${\mathrm{M}_\odot}$ object (see Chabrier & Bara ff e 2000). The evolution is calculated with our recent so-called ATMO atmosphere models and updated solar abundances (Ca ff au et al. 2011) for a global helium and heavy element abundance $Y=0.275, Z_\odot=0.017$, respectively (Phillips et al. 2020), but with 3 di ff erent H / He EOS, namely Saumon et al. (1995, SCvH), Chabrier et al. (2019, CMS19) and Chabrier & Debras (2021, CD21). The di ff erences between these di ff erent EOS, and the improvements in the treatment of the H-He interactions, are described in detail in these papers. As already mentioned in Phillips et al. (2020) and Chabrier & Debras (2021) and clearly seen in the figure, the new EOSs yield denser and cooler struc- tures for a given object, thus more correlated and degenerate (lower $\psi$) interiors. This in turn increases the cooling rate of the object. This is illustrated in the right panel of Fig. 2, which por- trays the late evolution of the e ff ective temperature, luminosity and radius of a 0.073 (black), 0.074 (blue) and 0.075 (red) ${\mathrm{M}_\odot}$ object for the same 3 EOSs. While these 3 quantities, notably $\mathrm{T_{eff}}$ and $L$, become constant after $\sim 5\times 10^8$ yr at $0.073\,{\mathrm{M}_\odot}$ with the SCvH EOS, indicating the stellar-substellar boundary, they keep decreasing with the two other EOSs, this limit occuring at $0.074\,{\mathrm{M}_\odot}$ with the CMS19 EOS and at $0.075\,{\mathrm{M}_\odot}$ with the CD21 EOS. Table 1 displays the characteristic of the steady H-burning limit (HBL) obtained with the 3 EOSs. We define the HBL as the limit below which nuclear equilibrium ($L_{\rm nuc}=L_\star$) will never be reached, thus below which cooling and gravitational contraction will last for ever. This corresponds to the physical limit of the stellar main sequence, i.e. the stellar-substellar boundary. Any object below this limit, i.e. cooler and fainter than the corre- sponding e ff ective temperature and luminosity (see Table 1) will be a brown dwarf, whatever its age. Note, however, that the op- posite is not true: objects hotter and brighter than these limits can be either stars or BDs, depending on their age. The H-burning minimum mass (HBMM), i.e. the minimum mass to sustain hy- drogen fusion, that does not depend on the age, is found to be 0.075 ${\mathrm{M}_\odot}$ ($\sim78.5\,{\mathrm{M_{Jup}}}$) with the most recent CD21 EOS. For sake of comparison, we also give in this table the characteris- tics of the HBL obtained in Bara ff e et al. (2003). These models use the SCvH EOS, the same helium and heavy element abun- dances as mentioned above but the so-called COND model at- mospheres. These comparisons enable us to disentangle the im- pact of the atmosphere and EOS models, respectively, upon the cooling properties of an object at the H-burning limit. The atmo-
221207153/6
Gilles Chabrier etal.:ImpactofanewH / Heequationofstateontheevolutionofmassivebrowndwarfs Gyr. The COND models (Bara ff e et al. 2003) for such temper- atures predict a mass $M=36\pm5\,{\mathrm{M_{Jup}}}$ for the BD, in strong conflict with the dynamical mass. Conversely, for $M=66\,{\mathrm{M_{Jup}}}$, these models predict $\mathrm{T_{eff}}\sim 1200\pm170$ K at the age of the sys- tem, significantly higher than the afore estimates from the best- fit spectral models. Table 2 displays the temperature, luminos- ity and radius obtained with the Phillips et al. (2020) models, which use the CMS19 models, and the present ones, with the CD21 EOS. Both models use the same abundances for helium, $Y=0.275$, and metals, $Z_\odot=0.017$, yielding an ’equivalent’ helium abundance $Y_{eq}=0.292$. This comparison highlights the impact of the H / He interactions in the most recent EOS mod- els. As seen in the table, even though, for the relevant mass and age range, the models including the new EOS yield temperatures cooler and luminosities fainter than the Bara ff e et al. (2003) and Phillips et al. (2020) models, relieving part of the tension, it is clear that they are still far from resolving the disagreement with the observational determinations. As noted by Cheetham et al. (2018), this discrepancy may be caused by the object being an unresolved binary brown dwarf system or by the presence of an additional object in the system, which could have biased the RV data and caused an overestimate of the dynamical mass. An equal mass binary of 500-600 K ob- jects with $R\sim 1\,{\mathrm{R_{Jup}}}$, notably, would provide a good match to the observed data while being in good agreement with the model predictions, namely $\mathrm{T_{eff}}=600\pm40$ K for a 33 ${\mathrm{M_{Jup}}}$ object at $5\pm1$ Gyr (e.g. Phillips et al. 2020 or present models). ## 3.2. Eps Indi Bab Recently, Chen et al. (2022) reported dynamical masses for the binary BD system $\epsilon$ Indi Ba (Sp = T1-1.5) and $\epsilon$ Indi Bb (Sp = T6, also called $\epsilon$ Indi C), with individual masses ${\mathrm{M_{dyn}}}=66.92\pm0.36\,{\mathrm{M_{Jup}}}$ and ${\mathrm{M_{dyn}}}=53.25\pm0.29\,{\mathrm{M_{Jup}}}$, respectively, with a $\approx 5\%$ precision. With an age of 3.5 $^{+0.8}_{-1.0}$ Gyr from $\epsilon$ Indi A’s activity, this system provides a stringent constraint for BD cooling models, notably for old, massive BDs. Field-aged ob- jects of spectral types T1 and T6 have e ff ective temperatures in the range $\mathrm{T_{eff}}\simeq 1300$ -900 K (e.g. Filipazzo et al. 2015). This is consistent with the inferred temperatures for Indi B and C, namely $\mathrm{T_{eff}}\simeq 1300$ -1340 K and $\mathrm{T_{eff}}\simeq 880$ -940 K, respectively (King et al. 2010). Correct theoretical evolutionary models must thus allow these objects to reach these spectral types and tem- peratures within the aforementioned timescale for the observed metallicity of the host star $\epsilon$ Indi A ($[Fe/H]=0.13$). nan of the widely used Chabrier et al. (2000), Burrows et al. (2001) and Saumon & Marley (2008) ’cloudy’ models can fullfill this con- straint and predict too high temperatures and luminosities. The ’COND’ Bara ff e et al. (2003) predict faster cooling rates, due to the less opaque atmospheres. However, as noted in § 2, the lower opacity in the COND atmosphere models stems, notably, from missing line opacities. Therefore, this analysis suggests that the aforementioned models underpredict the cooling rates for $\epsilon$ Indi Ba and Bb. Figure 4 displays the $L$ - $t$ and $L$ - $M$ relationships obtained for this system with the Phillips et al. (2020) cooling models, based on the CMS19 H / He EOS, and the present ones, based on the CD21 EOS, using the same ATMO atmosphere models and ef- fective helium + heavy element abundance $Y_{eff}=0.292$. As seen in the Figure, while the Phillips et al. (2020) models yield an age of between $\sim 3.0$ and 3.8 Gyr for the system for the observed lu- minosities, the present model predict slightly younger ages, of 2.8 Gyr for Ba and 3.5 Gyr for Bb. Both models are in good agreement with the observational determination. Assuming co- evality, this corresponds to a $\sim 0.7$ Gyr uncertainty on the age of these BDs, for this age and mass range. As seen in Fig. 13 of Chen et al. (2022), the Saumon & Marley (2008) hybrid mod- els predict a significantly older age (5 Gyr) for the system, at odds with the observational determination, suggesting a too slow cooling rate for these models. The corresponding e ff ective tem- peratures and surface gravities between 2.8 and 3.5 Gyr with the present models are $\mathrm{T_{eff}}=$ 1354-1265 K, $\log \,g=$ 5.45 for $\epsilon$ Indi Ba, for a mass 67 ${\mathrm{M_{Jup}}}$, and $\mathrm{T_{eff}}=$ 1060-998 K, $\log \,g=$ 5.34 for $\epsilon$ Indi Bb, for a mass 53 ${\mathrm{M_{Jup}}}$, respectively (see Table 2). As illustrated in Figure 4 and discussed earlier, the faster cooling rate for the most massive BDs with the CD21 EOS than with the CMS19 one stems the sharp increase of degeneracy in this mass range (see e.g. Fig. 1 of Chabrier & Bara ff e 2000). The cooling rate then appears to be a bit too slow as a function of the mass, with $\Delta \log L/\Delta \log M\simeq 3.95$ instead of 5.37 between $\epsilon$ Indi Ba and Bb, i.e. from early T to late T, even though the discrepancy remains within 2 $\sigma$. As mentioned earlier, this stems more likely from remaining issues with the atmosphere models than with the EOS. ## 3.3. Gl 758 B Combining radial velocity and astrometry, Bowler et al. (2018) determined a dynamical mass of ${\mathrm{M_{dyn}}}=42^{+19}_{-7}\,{\mathrm{M_{Jup}}}$ for the T7- T8 BD Gl 758 B, with a robust lower limit of 30.5 ${\mathrm{M_{Jup}}}$ at the 4 $\sigma$ level, for nominal ages 1-6 Gyr adopted for the host star (Vigan et al. 2016). More recently, using Hipparcos and Gaia data, Brandt et al. (2021) derived the most precise mass mea- surement to date for this system, with ${\mathrm{M_{dyn}}}=38.1^{+1.7}_{-1.5}\,{\mathrm{M_{Jup}}}$, while their analysis of activity and rotation of Gl 758A favors an age $\gtrsim$ 6 Gyr. As for the previous objects, substellar evolu- tionary models generally underestimate the mass of Gl 758 B. As noted by the authors above, this discrepancy can be rec- onciled if the system is older, which is consistent with activ- ity indicators and recent isochrone fitting of the host star, or alternatively if the models are systematically overluminous by $\approx$ 0.1-0.2 dex. Atmospheric model fitting yields for this object a bolometric luminosity $\log (L/L_\odot)=-6.07\pm0.03$, $\mathrm{T_{eff}}=650$ K and $\log g=5.0$. All current models essentially underpredict the mass for an age 1-6 Gyr or, alternatively, are overluminous by $>0.1$ dex at this age. This is in the opposite sense from results by Dupuy et al. (2009, 2014), who found that substellar cool- ing models underpredict the luminosities of brown dwarfs with dynamical masses by $\approx$ 0.2-0.4 dex. Altogether, the most likely explanation for the disagreement in mass probably resides in the age of Gl 758. Older ages of 6-9 Gyr would readily put the pre- dicted and dynamical distributions in excellent agreement and are indeed suggested from the low activity level, lack of X-ray emission and slow projected rotational velocity (see Bowler et al. (2018) and references therein). Indeed, more recent isochrone fitting are converging on an older value that agrees better with activity indicators, with an average of $5.3$ - $7.5$ Gyr (Brewer et al. 2016, Luck 2017). This is supported by the results displayed in Table 2, with the new models yielding a nearly perfect agree- ment with the observations for a mass $\sim 30$ - $40\,{\mathrm{M_{Jup}}}$ and an age $\sim$ 5-8 Gyr. ## 3.4. WISE system J072003.20 Individual dynamical masses for the nearby M9.5 + T5.5 binary system WISE J0720-0846AB have been determined by Dupuy et
221207153/4
Gilles Chabrier etal.:ImpactofanewH / Heequationofstateontheevolutionofmassivebrowndwarfs Fig.3. E ff ective temperature, luminosity and radius for $M\le0.075\,{\mathrm{M}_\odot}$ for 0.1, 1, 2, 5 and 10 Gyr (from top to bottom), respectively, for the same atmosphere models and helium and heavy element compositions but 2 di ff erent EOSs, namely Chabrier & Debras (2021) (present) and Chabrier et al. (2019) (Phillips et al. 2020). & Bara ff e 2000, Chabrier et al. 2009), at ${\rm T_{{eff}}}_{\rm HBL} \simeq 2075$ K, $\log(L/L_\odot)_{\rm HBL}\simeq -3.9$ and $R/R_\odot \simeq 0.086$. The inversion of the radius trend occurs near the location of the L2.5 dwarf 2MASS J0523-1403 ($M=67.54\pm12.79\,{\mathrm{M_{Jup}}}$, Sp = L2.5 (Filippazzo et al. 2015)). Although at first glance this seems to point to a dis- agreement with the models in the determination of the HBL, it must be kept in mind that the age of the objects in the Dieterich et al. (2014) sample, although supposedly larger than 1 Gyr, is unknown. This inevitably translates into some uncertainty in the comparison between observations and evolutionary calcula- tions. As seen in the right panel of Fig. 2, however, the HBL is only reached at much older ages, about 10 Gyr with the new EOS. Therefore, most of the objects identified in Dieterich et al. (2014), which lie at less than 25 pc and belong dominantly to the young disk, are too young to yield a proper determination of the HBL. On the other hand, as seen in figure 3, which por- trays the $\mathrm{T_{eff}}$ - $M$, $L$ - $M$ and $R$ - $M$ relations below the HBMM for 5 isochrones, an average age of about 2 Gyr for the observed sample near the $R$ - $\mathrm{T_{eff}}$ and $R$ - $L$ minimum location yields an ex- cellent agreement between the observational determinations and the new models. As highlighted in the rightmost panel of this figure, the minimum in the radius-mass relation does not occur exactly at the HBMM but at about 65 ${\mathrm{M_{Jup}}}$ (see, e.g., Fig. 1 of Chabrier et al. 2009), consistent with the aforementioned BD 2MASS J0523-1403 determination, and is more and more pro- nounced as objects get older. We see from the 3 panels that the observational identifications of the minimum radius in the sam- ple of Dieterich et al. (2014) are consistent with objects of mass $\sim 0.065$ - $0.075\,{\mathrm{M}_\odot}$ and age $\sim1$ -2 Gyr. We also see that, whereas the newest EOS a ff ects only modestly the e ff ective temperature and the luminosity, for a given mass and age, compared with the CMS19 EOS used in the models of Phillips et al. (2020), it a ff ects more severely the radius, notably for the most massive BDs. This reflects the di ff erent rate at which degeneracy pro- gresses throughout the body, starting at about $\sim 0.1$ Gyr, as seen in the left panel of Fig. 2. The Phillips et al. (2020) models have been updated accordingly and are available online (see at the end of the paper). It is important to mention at this stage that the procedure used by Dieterich et al. (2014) remains model atmosphere de- pendent, notably in a temperature domain characterised by cloud condensation and sedimentation in the models. Even though the BT-Settl atmosphere models, used by Dieterich et al. (2014), have been shown to provide a remarkable agreement with ob- served M and L spectra, alternative ’cloudless’ models have been proposed, which are very successfull in reproducing BD atmo- sphere spectra (Tremblin et al. 2015, 2016, 2017). ’Cloudless’ in these models does not mean that clouds do not form, but rather that they are not responsible for shaping the spectral evolution of brown dwarfs. Instead, the (massive) BD spectral evolution is due to a thermo-compositional instability triggered by the chem- ical conversion $CO\rightarrow CH_4$ in their atmospheres (Tremblin et al. 2019). Concerning the domain of very-low mass stars (VLM, $M\le 0.4\,{\mathrm{M}_\odot}$) on the main sequence (age $\gtrsim 1$ Gyr), the impact of the new EOS (CD21) compared with the SCvH one used in the Bara ff e et al. (2015, BHAC) models is quite modest and thus will not be displayed here. ## 3. Comparison with the data As mentioned in the previous section, near-infrared spectro- scopic and imaging surveys have uncovered a population of short-period, spectral binaries composed of low-mass stars and brown dwarfs, allowing a precise dynamical determination of the mass of these latter objects. Activity and kinematic constraints for the age of the primary of these massive T-dwarfs lead to a range of 2-8 Gyr. In this section, we compare the e ff ective tem- peratures obtained with the COND models (Bara ff e et al. 2003), the Phillips et al. (2020) models, based on the Chabrier et al. (2019) EOS, and a subset of models based on the most recent Chabrier & Debras (2021) EOS, with the various observations of BD dynamical systems. ## 3.1. HD4113C
221207153/1
# Impact of a new H/He equation of state on the evolution of massive brown dwarfs. New determination of the hydrogen burning limit Gilles Chabrier 1, 2, Isabelle Bara ff e 2, 1, Mark Phillips 3, 2,and Florian Debras 4 We have explored the impact of the latest equation of state (EOS) for dense hydrogen-helium mixtures (Chabrier & Debras 2021), which takes into account the interactions between hydrogen and helium species, upon the evolution of very low mass stars and brown dwarfs (BD). These interactions modify the thermodynamic properties of the H / He mixture, notably the entropy, a quantity of prime importance for these fully convective bodies, but also the onset and the development of degeneracy throughout the body. This translates into a faster cooling rate, i.e. cooler isentropes for a given mass and age, and thus larger brown dwarf masses and smaller radii for given e ff ective temperature and luminosity than the models based on previous EOSs. This means that objects of a given mass and age, in the range $M\lesssim 0.1\,{\mathrm{M}_\odot}$, $\tau\gtrsim 10^8$ yr, will have cooler e ff ective temperatures and fainter luminosities. Confronting these new models with several observationally determined BD dynamical masses, we show that this improves the agreement between evolutionary models and observations and resolves at least part of the observed discrepancies between the properties of dynamical mass determinations and evolutionary models. A noticeable consequence of this improvement of the dense H / He EOS is that it yields a larger H-burning minimum mass, now found to be $0.075\,{\mathrm{M}_\odot}$ ($78.5\,{\mathrm{M_{Jup}}}$) with the ATMO atmosphere models for solar metallicity. These updated brown dwarf models are made publicly available. ## 1. Introduction Tremendous progress has been accomplished along the years in the theoretical description of brown dwarfs (BD), enabling us to better understand their fundamental properties and cooling his- tories. The most recent progresses include a more complete de- scription of their atmosphere, thus of their spectral energy distri- bution (e.g. Morley et al. 2012, 2014, Tremblin et al. 2015, 2016, 2017), and of their interior, notably concerning the equation of state (EOS) of dense hydrogen and helium (Chabrier et al. 2019, CMS19). A new generation of BD evolutionary models has been derived recently, that incorporates both new (so-called ’ATMO’) atmosphere models and the new CMS19 H / He EOS (Phillips et al. 2020). As shown in these models, a noticeable impact of this new EOS is to yield denser and cooler, thus more degener- ate objects at the same mass than the ones computed with the Saumon-Chabrier-vanHorn (SCvH) EOS (Saumon et al. 1995). This yields slightly faster cooling rates, thus cooler temperatures and lower brightness at a given age (Phillips et al. 2020). Yet, a puzzling issue has emerged within the recent years, thanks to the determination of the dynamical masses of several ’massive’ T dwarfs (Cheetham et al. 2018, Dieterich et al. 2018, Dupuy et al. 2019, Bowler et al. 2018, Sahlmann et al. 2020, Brandt et al. 2019). These observations reveal dynamical masses significantly larger than the theoretical predictions for the deter- mined e ff ective temperature and age. A common feature of all these objects is their relatively high mass ($\sim60$ -75 ${\mathrm{M_{Jup}}}$), near the hydrogen burning minimum mass (HBMM), similar spec- tral type (late T) and cool e ff ective temperatures ($\mathrm{T_{eff}}\lesssim 1200$ K), which suggests ages in the range $\sim 5$ -10 Gyr, at the very least $>1$ Gyr. A noticeable exemple of this puzzling issue is the recent dynamical mass determination of Gliese 229 B, with a larger mass than estimated previously (Brandt et al. 2020). For all these late, massive T dwarfs, all existing BD evolutionary models generally underpredict the mass for the nominal ages and temperatures, suggesting too small cooling rates for these objects. This trend is confirmed by the recent thorough analysis of BD companion dynamical masses using Hipparcos and Gaia EDR3 data and age determinations based on activity-rotation- age calibrations by Brandt et al. (2021). For old ($\gtrsim$ 5 Gyr) and high-mass ($\gtrsim 60\,{\mathrm{M_{Jup}}}$) BDs, the models overpredict luminosities for the measured mass and age, or equivalently underestimate (resp. overestimate) the mass (resp. the age) for the correct age (resp. mass). As noticed by Dieterich et al. (2018), concerning the Eps Indi B-C system, models cannot make such massive ob- jects reach such a cool $\mathrm{T_{eff}}$ within the age of the Galaxy. Due to the degeneracy between mass and age in the substellar domain, there are no reliable age indicators for isolated BDs, and thus no robust constraint can be derived regarding substellar cooling rates and evolutionary models. In contrast, the aforementioned BD companions to higher mass stars with masses close to the stellar-substellar limit allow us to test a boundary value of the theory of substellar structure and evolution. A related diagnostic of these observations is the suggestion that the HBMM, identified as a minimum in the radius-e ff ective temperature and radius-luminosity relation, is larger and lies at a $\sim 400$ K larger e ff ective temperature than predicted by all cur- rent models (Dieterich et al. 2014). In this paper, we show that the most recent improvements in EOS calculations of dense hydrogen / helium mixtures contribute to resolve these issues in BD cooling theory. From a more gen-
221207153/3
Gilles Chabrier etal.:ImpactofanewH / Heequationofstateontheevolutionofmassivebrowndwarfs Fig.2. Left: Evolution of the central density, temperature and degeneracy parameter for a 0.075 ${\mathrm{M}_\odot}$ object calculated with 3 dif- ferent EOS, namely CD2021 (solid line), CMS2019 (long-dashed line) and SCvH (dotted line). Right: Evolution of the e ff ective temperature and luminosity for 0.075 (red), 0.074 (blue) and 0.073 (black) ${\mathrm{M}_\odot}$ and radius for 0.075 ${\mathrm{M}_\odot}$, with the same 3 EOS. \begin{tabular}{lcccc} \hline\hline EOS & ${\mathrm{M}_{\rm HBMM}}/M_\odot$ & ${\rm T_{{eff}}}_{\rm HBL}$ [K] & $\log(L/L_\odot)_{\rm HBL}$ & $R_{\rm HBL}/R_\odot$ \hline SCvH+COND (Baraffe et al. 2003) & 0.072 & 1560 & -4.47 & 0.081 SCvH+ATMO & 0.073 & 1807 & -4.14 & 0.087 CMS'19+ATMO (Phillips et al. 2020) & 0.074 & 1800 & -4.16 & 0.085 CD'21+ATMO (present) & 0.075 & 1800 & -4.19 & 0.083\hline \end{tabular} Table1. Mass, e ff ective temperature, luminosity and radius characteristic of the H-burning limit ($t$ = 10 Gyr) with the SCvH (1995), Chabrier et al. (2019) and Chabrier & Debras (2021) EOSs, all calculated with the ATMO atmosphere models. For the sake of comparison, the results with the COND atmosphere models are also presented. sphere models have been greatly improved between the COND and ATMO models. Line opacities, in particular, were missing in the former ones, yielding a much faster cooling rate and thus explaining the cooler and fainter limits at the HBL. But the EOS also bears some impact on the HBL. As seen in Table 1, the larger cooling rate with the new EOS translates into a larger HBMM with a smaller radius than in the previous calculations. On the other hand, the luminosities at the HBL remain barely a ff ected ($\lesssim 10\%$) and the e ff ective temperature remains essen- tially the same. This is explained by the fact that the threshold for hydrogen fusion at the center of the star occurs at a fixed temperature and thus, for the same interior-atmosphere (T $_c$ - $\mathrm{T_{eff}}$) boundary condition, at the same e ff ective temperature. In con- trast, the nuclear energy rate, thus the luminosity depends on the mass and then slightly di ff ers between the 3 di ff erent HBMM’s. On the other hand, Dieterich et al. (2014) have determined the e ff ective temperature and bolometric fluxes of 63 objects ranging in spectral type from M6V to L4, bracketing the stel- lar / substellar boundary, by comparing observed optical and in- frared photometric colors on 9 bands with synthetic ones, de- rived from the so-called BT-Settl model atmospheres (Allard et al. 2012, 2013), used in the Bara ff e et al. (2015) evolutionary calculations. The optimisation procedure for determining e ff ec- tive temperatures also indicates which model spectrum in the BT-Settl grid provides the overall best fit to the observed pho- tometry. Once the e ff ective temperatures and bolometric fluxes of the objects were determined by this procedure, the radii of the objects with known trigonometric parallax were determined from the Stefan-Boltzmann law (see Dieterich et al. (2014) for details). As the di ff erences between stellar and substellar objects become more pronounced at ages $>1$ Gyr, objects with known youth signatures were rejected from the sample. Based on this analysis, Dieterich et al. (2014) find evidence for a local min- imum in the radius-temperature and radius-luminosity trends, that indicates the vicinity of the stellar main sequence - brown dwarf sequence boundary (see e.g. Burrows et al. 1997, Chabrier
221207153/8
Gilles Chabrier etal.:ImpactofanewH / Heequationofstateontheevolutionofmassivebrowndwarfs find masses ${\mathrm{M_{dyn}}}=66^{+5}_{-14}\,{\mathrm{M_{Jup}}}$ and spectral type T5.5 for 2M0805 + 48 B and ${\mathrm{M_{dyn}}}=67^{+4}_{-5}\,{\mathrm{M_{Jup}}}$, T3.5 for 2M1059-21 B. The striking feature of their analysis is that the mass for the T3.5 2M1059-21B object is significantly higher than its two spectral type equivalents DENIS J2252-1730B (T3.5), 41 $\pm 4\, {\mathrm{M_{Jup}}}$, and 2MASS J1534-2952A (T4.5), 51 $\pm 5\, {\mathrm{M_{Jup}}}$. As noted by the authors, the mass of 2M0805 + 48B is almost equal to the one of WI0720-08B estimated by Dupuy et al. (2019), which has the same spectral type, yet a mass higher than the other three T5 dwarfs with measured dynamical masses, which all have masses lower than $<61\,{\mathrm{M_{Jup}}}$ as an upper limit (see Table 6 of Dupuy et al. 2019). As shown by the authors (see their Fig. 13), the masses derived for each member of our two pairs are compatible at the 1 $\sigma$ level with the 5 to 12 Gyr isochrones of Bara ff e et al. (2015), while $>$ 1 Gyr isochrones show reasonable agreement with the observational data. This, as the other previous analysis, highlights the fact that the higher observational masses than other T-dwarfs of similar spectral type and than predicted by the models, suggesting that models underpredict the mass for a given temperature or lumi- nosity, concerns essentially the most massive ($\gtrsim 60\,{\mathrm{M_{Jup}}}$) BDs, old enough to have reach the T spectral type domain, i.e. older than $\gtrsim 1$ Gyr for such masses. This strengthens our suggestion that this property stems essentially from the higher degeneracy, thus the faster cooling of these objects, which culminates at the highest central density of the stellar-substellar domain (see Fig. 1 of Chabrier & Bara ff e 2000). The new models predict cooler temperatures and fainter luminosities for a given mass and age compared with the Bara ff e et al. (2003) models, by up to about 100-150 K and $\sim 0.2$ - $0.3$ dex for the present ones, resolving at least partly the discrepancies mentioned above (see Table 2). Here again, we note that the mass and age of this system are similar to the ones of the BD HD19467B listed in Brandt et al. (2021), with an excellent agreement for the models. ## 3.6. Gliese 229B Combining Keck / HIRES radial velocities, imaging with HiCIAO / Subaru and the HST, and absolute astrometry from Hipparcos and GAIA, Brandt et al. (2020) measured the dynam- ical mass ${\mathrm{M_{dyn}}}=70\pm 5\,{\mathrm{M_{Jup}}}$ for the T7 BD Gliese 229B. Not only this value is higher than the $\simeq 64.0^{+2}_{-1}\,{\mathrm{M_{Jup}}}$ predicted by the Bara ff e et al. (2003) or Saumon & Marley (2008) models for such a low luminosity, $\log(L/L_\odot)=-5.208$, but to be com- patible with the observational determinations, the models would predict an age of 7-10 Gyr. Such an age seems to be excluded by kinematic and activity indicators that rather suggest a range 2-6 Gyr. Gliese 229B thus joins the club of ultracool BDs near the HBMM too massive, for their age, for the model predictions. For a $70\,{\mathrm{M_{Jup}}}$ at an age of 10 Gyr, the Phillips et al. (2020) mod- els (see their Fig. 8), calculated with the Chabrier et al. (2019) EOS, are $\sim 0.1$ dex less luminous than the Bara ff e et al. (2003, B03) and $\sim 0.4.$ dex less luminous than the hybrid cloud tracks of SM08, helping to relieve some of the tension. As seen in Table 2, the present models, based on the new CD21 EOS, predict $\gtrsim$ 40 K cooler and $\sim -0.1$ dex fainter models than the Phillips et al. (2020) ones for a mass and luminosity consistent with the lat- est observationally inferred values, for an age of about 9-10 Gyr, improving the model-observations agreement. Best-fit atmosphere models to the observed spectrum of Gliese 229B by Nakajima et al. (2015) yield acceptable solutions in the range 750 K $\le\mathrm{T_{eff}}\le$ 900 K, $4.75\le \log g\le 5.0$. As seen in the Table, these values are quite small compared to the ones from the models. It must be stressed, however, that this fitting proce- dure is based on one single source of cloudy atmosphere models, namely Tsuji (2002, 2005). A more robust determination of the e ff ective temperature and gravity requires further detailed com- parisons with more recent atmosphere models, a point we have already stressed in § 2. As a last remark, it is worth noting that a possibility to resolve the discrepancy between models and observations for Gl229 B would be that this latter is itself an unresolved tight binary (see e.g. Brandt et al. 2021). A case close to Gliese 229B is HR 7672 B, with a dy- namical mass ${\mathrm{M_{dyn}}}=72.7\pm{0.8}\,{\mathrm{M_{Jup}}}$ and a weighted aver- age luminosity $\log(L/L_\odot)=-4.25\pm 0.05$ (Brandt et al. 2021), while the activity analysis suggests a rather young age, centered around 2 Gyr (Brandt et al. 2019, 2021). For this age and a mass $M=70\,{\mathrm{M}_\odot}$ ($=72.7\,{\mathrm{M_{Jup}}}$), the present models predict $\log(L/L_\odot)=-4.25$, $\mathrm{T_{eff}}=$ 1730 K, while the Phillips et al. (2020) ones yield $\log(L/L_\odot)=-4.16$, $\mathrm{T_{eff}}=$ 1785 K. In order to illustrate the results listed in Table 2, Fig. 5 dis- plays the $\mathrm{T_{eff}}$ -mass and $L$ -mass data comparisons (as in Fig. 8 of Phillips et al. 2020), including the most recent determinations, for massive BDs for 4 isochrones, namely 1, 2, 5 and 8 Gyr, based on the CMS19 (Phillips et al. 2020) and CD21 (present) EOS models. To briefly summarise this section, we note that the models based on the most recent EOSs, notably the latest CD21 one, reach cooler temperatures and fainter luminosities than the pre- vious generation for a given mass and age, illustrating the faster cooling rates for massive, old ($\gtrsim 1$ Gyr) BDs, yielding in gen- eral a much better agreement with the observational determina- tion. As seen in Table 2, some tension remains potentially for Gl229B but most importantly for HD4113C, which appears to be substantially cooler and fainter than predicted by the mod- els. Given the general agreement for all the other objects, this suggests that these two BDs could be unresolved systems. ## 4. Conclusion In this paper, we have explored the impact of the latest EOS for dense hydrogen-helium mixtures upon the structure and evo- lution of very low mass stars and brown dwarfs. Whereas the previously used Saumon et al. (1995) and Chabrier et al. (2019) EOSs are based on the so-called additive volume law, which im- plies that the interactions between hydrogen and helium species are not taken into account, these interactions are included in the Chabrier & Debras (2021) EOS, based on the QMD simulations of Militzer & Hubbard (2013) (see CD21 for details). These interactions modify the thermodynamic properties of the H / He mixture in two ways. They yield cooler and denser, thus more compact structures, i.e. smaller radii and thus fainter luminosity for given mass and age. They also yield cooler entropy profiles, a quantity of prime importance for these fully convective bodies, which a ff ects the onset and the development of electron degen- eracy throughout the body (see Fig. 9 of CD21). This translates into a faster cooling rate and thus larger masses for given e ff ec- tive temperatures and luminosities at a given age in the brown dwarf regime (see Table 2). Confronting these new models with several observationally determined BD dynamical masses, we show that, indeed, this improves the agreement between evolu- tionary models and observations and resolves at least part of the observed discrepancies for massive, rather old brown dwarfs. In
200211922/5
Table2. Best-fitting parameters for Model 1: crabcor*TBabs (diskbb + relxill) \begin{tabular}{llccccccc} \rightarrowprule Model & Parameter &Spec. A & Spec. B & Spec. C & Spec. D & Spec. E & Spec. F & Spec. G \midrule {\tt crabcor}&$C$ &\multicolumn{7}{c}{1.097 (f)}\specialrule{0em}{1pt}{1.1pt} &$\Delta\Gamma$ &\multicolumn{7}{c}{0.01 (f)}\specialrule{0em}{1pt}{1.1pt} {\tt TBabs}&$N_\mathrm{H}$ (cm$^{-2}$) &\multicolumn{7}{c}{$4.0\times10^{21}$ (f)}\specialrule{0em}{1pt}{1.1pt} {\tt diskbb} &$T_\mathrm{col}$ (keV)&$0.858_{-0.014}^{+0.011}$&$0.825\pm0.012$&$0.835_{-0.018}^{+0.015}$&$0.787_{-0.010}^{+0.008}$&$0.787_{-0.009}^{+0.007}$&$0.771\pm0.007$&$0.772_{-0.009}^{+0.010}$\specialrule{0em}{1pt}{1.1pt} &$N_\mathrm{DISC}$&$3889_{-219}^{+366}$&$4151_{-273}^{+237}$&$3561_{-297}^{+312}$&$4862_{-247}^{+329}$&$4923_{-234}^{+398}$&$5025_{-212}^{+230}$&$4891_{-311}^{+328}$\specialrule{0em}{1pt}{1.1pt} {\tt relxill} &$q$&$3.69_{-0.56}^{+1.03}$&$3.78_{-0.64}^{+1.23}$&$3.29_{-0.52}^{+1.04}$&$3.98_{-0.70}^{+1.41}$&$3.82_{-0.57}^{+1.09}$&$3.67_{-0.55}^{+1.20}$&$3.71_{-0.65}^{+1.13}$ \specialrule{0em}{1pt}{1.1pt} &$a_*$&\multicolumn{7}{c}{$0.67_{-0.08}^{+0.15}$}\specialrule{0em}{1pt}{1.1pt} &$i$ (deg)&\multicolumn{7}{c}{$36.3_{-3.4}^{+5.3}$}\specialrule{0em}{1pt}{1.1pt} &$A_\mathrm{Fe}$&\multicolumn{7}{c}{$5.05_{-0.26}^{+1.21}$}\specialrule{0em}{1pt}{1.1pt} &$\Gamma_\mathrm{r}$&$2.79\pm0.06$&$2.59_{-0.04}^{+0.06}$&$2.65_{-0.07}^{+0.08}$&$2.71_{-0.09}^{+0.06}$&$2.66_{-0.08}^{+0.07}$&$2.69_{-0.08}^{+0.06}$&$2.80_{-0.07}^{+0.12}$ \specialrule{0em}{1pt}{1.1pt} &$\mathrm{log}\xi$&$3.72_{-0.16}^{+0.13}$&$3.54_{-0.13}^{+0.22}$&$3.53_{-0.15}^{+0.46}$&$3.70_{-0.21}^{+0.13}$&$3.72_{-0.19}^{+0.29}$&$3.73_{-0.12}^{+0.15}$&$3.71_{-0.18}^{+0.05}$\specialrule{0em}{1pt}{1.1pt} &$R_\mathrm{f}$&$0.35_{-0.10}^{+0.07}$&$0.29_{-0.03}^{+0.06}$&$0.27_{-0.04}^{+0.07}$&$0.44_{-0.11}^{+0.05}$&$0.41_{-0.08}^{+0.06}$&$0.44_{-0.07}^{+0.16}$&$0.61_{-0.14}^{+0.12}$ \specialrule{0em}{1pt}{1.1pt} &$N_\mathrm{r}$&$0.184_{-0.031}^{+0.038}$&$0.091_{-0.011}^{+0.016}$&$0.099_{-0.022}^{+0.024}$&$0.065_{-0.016}^{+0.013}$&$0.051_{-0.012}^{+0.011}$&$0.04_{-0.009}^{+0.008}$&$0.057_{-0.012}^{+0.026}$ \midrule & $\chi^2$ &\multicolumn{7}{c}{390.40} & $\nu$ & \multicolumn{7}{c}{452} & $\chi^2_{\nu}$ & \multicolumn{7}{c}{0.86} \bottomrule \end{tabular} Notes. Columns 3-9 show successively the results of Spec. A-G. The parameters with “ f ” in parenthesis indicate they were fixed at values given. All errors for one parameter of interest were calculated with 90% confidence level. 1. Parameters including the spin $a_*$, the inclination angle $i$ and the iron abundance $A_\mathrm{Fe}$ of relxill were linked together among different spectra. 2. Parameters including the temperature $T_\mathrm{col}$, normalization constant $N_\mathrm{DISC}$, emissivity index $q$, photon index $\Gamma _\mathrm{r}$, ionization state $\mathrm{log}\xi$, reflection fraction $R_\mathrm{f}$ and the normalization $N_\mathrm{r}$ were independent for each spectrum. upper limit, which indicates that the density in the disc is larger than the maximal value in the model. We found that the spin parameter pegged at -0.998 when the iron abundance was decreased to $\sim{3}$. To better understand this surprising finding, we did another trial. We fixed the iron abundance at values between 3.5-6.5 with a stepsize of 1.0, and found four best fits with $\chi^2_{\nu}$ < 1 (d.o.f = 453). We, then, explored the $\chi^2$ for spins from 0 to 1.0 with stepsize of 0.01 for them (Figure 7). These four models all obtain moderate spin black holes at 90% statistical confidence level, but when the iron abundance is higher than $\sim{5.5}$, at more than 90% statistical confidence level, we note the reduced sensitivity of models to large values of spin. The increase of iron abundance induces more photoelectric absorption making the Fe K-edge near 8 keV deeper. At the same time, the strength of the Fe k emission in the band of 6-8 keV increases (Garcíaetal.2013). More fluorescent iron photons near the black hole would be scattered down below $\sim{6}$ keV to make a stronger red wing, as expected. Then, we further explored the dependence of the spin parameter $a_*$ on the inclination angle $i$. We fit spectra for 60 evenly-spaced values of the $a_*$ in the range of 0.4-1.0 and 30 evenly-spaced values of the $i$ with the range of 20 $^{\circ}$ -50 $^{\circ}$ (Figure 8). When $i$ is larger than $\sim{36}^{\circ}$, a positive relationship is shown. Moreover, the model lost the ability to give an upper limit on the spin parameter at 99% confidence level. The inclination angle ($36.3_{-3.4}^{+5.3}$ degrees) in this paper is consis- tent with the value ($32_{-4}^{+3}$ degrees) in, which may indicate that the inclination angle of the inner disc is misaligned with the orbital inclination angle ($\sim{21}^{\circ}$). However, for a transient system, the timescale for accretion to torque the black hole into alignment is approximated to be $10^{6}-10^{8}$ years (Martin etal.2008). Therefore, the alignment is expected to occur early in the typical lifetime of transients, which are characteristically Gyrs old (White&Ghosh1998; Fragosetal.2013). In our estimation, the most likely resolution to this apparent tension lies in the reflection modeling. The inclination angle estimation via X-ray reflection fit- ting method is principally determined by the blue wing of the broad Fe line. The high density model leads to increasing soft X-ray flux. Recent reflection analyses of Cyg X-1 by Tomsicketal. (2018), and GX 339-4 by Garcíaetal. (2015) and Jiangetal. (2019) suggest that reflection models which underestimate the density of disc introduce systematic changes of order 10 $\degr$ in the inclination angle. Shafeeetal. (2006) first reported the spin of the black hole in 4U 1543 via the continuum-fitting method. They estimated its spin to be $0.8\pm0.1$. Then, and Morningstar&Miller (reported two spin measurements, $0.3\pm0.1$ and $0.43^{+0.22}_{-0.31}$, respectively, both constrained by combining the continuum-fitting and X-ray reflection fitting methods. These three works utilized the dynamical parameters which were reported in Park et al. (2004), but found conflicting values of spin. For measuring the spin of a black hole via the continuum-fitting method to succeed, one con- strains the size of the emitting region via the efficient blackbody-like property of the optically thick disc. To relate the emitting area to a dimensionless ISCO and thereby spin, it is critical to have accurate measurements of the distance to the source, the mass of the black hole, and the inclination angle of the accretion disc (McClintock etal.2011; McClintocketal.2006; Gouetal.2011). The spin and the inclination angle measured by modeling the reflection emission with relxill in this paper is consistent with those reported by Morningstar&Miller (2014). However, the spin measurement is in conflict with the one reported by Miller et al. (2009). Milleretal. (2009) assumed the inclination angle of accre- tion disc is equal to the orbital inclination angle. Accordingly, we test the implication of the lower inclination angle on the spin measurement. When the inclination angle param- eter in REXILL model was fixed at the orbital inclination angle of 21.0 $^{\circ}$, which is considered as Model 2, we find that the fit be-
200211922/7
Table3. Best-fitting parameters for Model 2: crabcor*TBabs *(diskbb + relxill), in which $i$ = 21 deg \begin{tabular}{llccccccc} \rightarrowprule Model & Parameter &Spec. A & Spec. B & Spec. C & Spec. D & Spec. E & Spec. F & Spec. G \midrule {\tt crabcor}&$C$ &\multicolumn{7}{c}{1.097 (f)}\specialrule{0em}{1pt}{1.1pt} &$\Delta\Gamma$ &\multicolumn{7}{c}{0.01 (f)}\specialrule{0em}{1pt}{1.1pt} {\tt TBabs}&$N_\mathrm{H}$ ($cm^{-2}$) &\multicolumn{7}{c}{$4.0\times10^{21}$ (f)}\specialrule{0em}{1pt}{1.1pt} {\tt diskbb} &$T_\mathrm{col}$(keV)&$0.842_{-0.013}^{+0.015}$&$0.814_{-0.015}^{+0.016}$&$0.829\pm0.015$&$0.784\pm0.013$&$0.783\pm0.007$&$0.770_{-0.007}^{+0.008}$&$0.770_{-0.011}^{+0.010}$ \specialrule{0em}{1pt}{1.1pt} &$N_\mathrm{DISC}$&$4349_{-291}^{+220}$&$4471_{-374}^{+376}$&$3778_{-301}^{+229}$&$5021_{-383}^{+388}$&$5124_{-216}^{+221}$&$5110_{-272}^{+251}$&$5052_{-306}^{+355}$ \specialrule{0em}{1pt}{1.1pt} {\tt relxill} &$q$&$2.72_{-0.13}^{+0.10}$&$2.69\pm0.13$&$2.44\pm0.17$&$2.78_{-0.14}^{+0.08}$&$2.67\pm0.09$&$2.59_{-0.06}^{+0.11}$&$2.76_{-0.15}^{+0.10}$\specialrule{0em}{1pt}{1.1pt} &$a_*$&\multicolumn{7}{c}{$> 0.83$}\specialrule{0em}{1pt}{1.1pt} &$i$ (deg)&\multicolumn{7}{c}{$21.0$ (f)}\specialrule{0em}{1pt}{1.1pt} &$A_\mathrm{Fe}$&\multicolumn{7}{c}{$7.72_{-1.62}^{+1.23}$}\specialrule{0em}{1pt}{1.1pt} &$\Gamma_\mathrm{r}$&$2.64_{-0.03}^{+0.04}$&$2.48_{-0.03}^{+0.04}$&$2.51_{-0.03}^{+0.04}$&$2.56_{-0.04}^{+0.09}$&$2.52\pm0.03$&$2.53_{-0.02}^{+0.04}$&$2.59_{-0.05}^{+0.14}$ \specialrule{0em}{1pt}{1.1pt} &$\mathrm{log}\xi$&$4.52_{-0.15}^{+4.52}$&$4.19_{-0.35}^{+0.25}$&$4.26_{-0.34}^{+0.15}$&$4.07_{-0.34}^{+0.25}$&$4.29_{-0.20}^{+0.09}$&$4.19_{-0.23}^{+0.16}$&$4.01_{-0.23}^{+0.35}$\specialrule{0em}{1pt}{1.1pt} &$R_\mathrm{r}$&$0.21_{-0.09}^{+0.06}$&$0.23_{-0.01}^{+0.06}$&$0.18_{-0.03}^{+0.04}$&$0.3_{-0.04}^{+0.06}$&$0.32\pm0.04$&$0.32_{-0.04}^{+0.05}$&$0.36_{-0.06}^{+0.13}$ \specialrule{0em}{1pt}{1.1pt} &$N_\mathrm{r}$&$0.111_{-0.011}^{+0.015}$&$0.063_{-0.008}^{+0.010}$&$0.064_{-0.007}^{+0.010}$&$0.04_{-0.003}^{+0.015}$&$0.032_{-0.003}^{+0.004}$&$0.023_{-0.003}^{+0.004}$&$0.031_{-0.005}^{+0.016}$ \midrule & $\chi^2$ &\multicolumn{7}{c}{443.71} & $\nu$ & \multicolumn{7}{c}{453} & $\chi^2_{\nu}$ & \multicolumn{7}{c}{0.98} \bottomrule \end{tabular} Notes. Columns 3-9 show successively the results of Spec. A-G. The parameters with “ f ” in parenthesis indicate they were fixed at values given. All errors for one parameter of interest were calculated with 90% confidence level. 1. Parameters including the spin $a_*$ and the iron abundance $A_\mathrm{Fe}$ of relxill were linked together among different spectra. 2. Parameters including the temperature $T_\mathrm{col}$, normalization constant $N_\mathrm{DISC}$, emissivity index $q$, photon index $\Gamma _\mathrm{r}$, ionization state $\mathrm{log}\xi$, reflection fraction $R_\mathrm{r}$ and the normalization $N_\mathrm{r}$ were independent for each spectrum. Table4. Best-fitting parameters for Model 3: crabcor*TBabs *(diskbb + relxill), in which $i$ = 21 deg and $a_{*}$ = 0.30 \begin{tabular}{llccccccc} \rightarrowprule Model & Parameter &Spec. A & Spec. B & Spec. C & Spec. D & Spec. E & Spec. F & Spec. G \midrule {\tt crabcor}&$C$ &\multicolumn{7}{c}{1.097 (f)}\specialrule{0em}{1pt}{1.1pt} &$\Delta\Gamma$ &\multicolumn{7}{c}{0.01 (f)}\specialrule{0em}{1pt}{1.1pt} {\tt TBabs}&$N_\mathrm{H}$ ($cm^{-2}$) &\multicolumn{7}{c}{$4.0\times10^{21}$(f)}\specialrule{0em}{1pt}{1.1pt} {\tt diskbb} &$T_\mathrm{col}$(keV)&$0.853\pm0.013$&$0.83\pm0.015$&$0.837_{-0.016}^{+0.013}$&$0.798\pm0.009$&$0.791_{-0.008}^{+0.007}$&$0.778_{-0.007}^{+0.006}$&$0.783_{-0.010}^{+0.008}$\specialrule{0em}{1pt}{1.1pt} &$N_\mathrm{DISC}$&$4127_{-172}^{+262}$&$4103_{-341}^{+346}$&$3586_{-242}^{+331}$&$4595_{-259}^{+324}$&$4868_{-212}^{+250}$&$4862_{-192}^{+239}$&$4603_{-241}^{+360}$\specialrule{0em}{1pt}{1.1pt} {\tt relxill} &$q$&$2.75\pm0.15$&$2.75_{-0.15}^{+0.07}$&$2.49_{-0.20}^{+0.18}$&$2.76_{-0.19}^{+0.15}$&$2.72_{-0.11}^{+0.10}$&$2.61_{-0.13}^{+0.12}$&$2.65\pm0.23$\specialrule{0em}{1pt}{1.1pt} &$a_*$&\multicolumn{7}{c}{$0.30$ (f)}\specialrule{0em}{1pt}{1.1pt} &$i$ (deg)&\multicolumn{7}{c}{$21.0$ (f)}\specialrule{0em}{1pt}{1.1pt} &$A_\mathrm{Fe}$&\multicolumn{7}{c}{$4.5_{-0.21}^{+0.46}$ }\specialrule{0em}{1pt}{1.1pt} &$\Gamma_\mathrm{r}$&$2.65\pm0.03$&$2.5_{-0.03}^{+0.05}$&$2.54\pm0.03$&$2.62_{-0.06}^{+0.08}$&$2.54\pm0.02$&$2.56\pm0.03$&$2.73_{-0.12}^{+0.06}$ \specialrule{0em}{1pt}{1.1pt} &$\mathrm{log}\xi$&$4.31_{-0.63}^{+0.22}$&$3.89_{-0.36}^{+0.29}$&$4.03_{-0.27}^{+0.30}$&$3.73_{-0.21}^{+0.31}$&$4.06_{-0.18}^{+0.24}$&$4.0_{-0.20}^{+0.17}$&$3.63_{-0.16}^{+0.27}$\specialrule{0em}{1pt}{1.1pt} &$R_\mathrm{r}$&$0.18\pm0.03$&$0.19_{-0.02}^{+0.03}$&$0.17\pm0.02$&$0.29\pm0.06$&$0.27_{-0.02}^{+0.04}$&$0.29\pm0.03$&$0.39_{-0.06}^{+0.09}$ \specialrule{0em}{1pt}{1.1pt} &$N_\mathrm{r}$&$0.115_{-0.012}^{+0.025}$&$0.069_{-0.008}^{+0.012}$&$0.07_{-0.008}^{+0.009}$&$0.049_{-0.009}^{+0.013}$&$0.034\pm0.003$&$0.026\pm0.003$&$0.046_{-0.014}^{+0.009}$ \midrule & $\chi^2$ &\multicolumn{7}{c}{456.69} & $\nu$ & \multicolumn{7}{c}{454} & $\chi^2_{\nu}$ & \multicolumn{7}{c}{1.01} \bottomrule \end{tabular} Notes. Columns 3-9 show successively the results of Spec. A-G. The parameters with “ f ” in parenthesis indicate they were fixed at values given. All errors for one parameter of interest were calculated with 90% confidence level. 1. The iron abundance $A_\mathrm{Fe}$ of relxill were linked together among different spectra. 2. Parameters including the temperature $T_\mathrm{col}$, normalization constant $N_\mathrm{DISC}$, emissivity index $q$, photon index $\Gamma _\mathrm{r}$, ionization state $\mathrm{log}\xi$, reflection fraction $R_\mathrm{r}$ and the normalization $N_\mathrm{r}$ were independent for each spectrum. ## ACKNOWLEDGEMENTS We thank the useful discussion with Prof. J.Orosz, Prof. Youjun Lu, Dr. Erlin Qiao, Dr. Weiwei Xu, and Dr. Zhu Liu. We would also like to thank the reviewer for his/her valuable input. Lijun Gou are supported by the National Program on Key Research and Development Project through grant No. 2016YFA0400804, and by the National Natural Science Foundation of China with grant No. U1838114, and by the Strategic Priority Research Program of the Chinese Academy of Sciences through grant No. XDB23040100. We also thank RXTE/PCA public data and facilities. This work is made under the help with tools available on Astrophysics Sci- ence Archive Research Centre (HEASARC), belonging to NASA’s Coddard Space Flight Centre (GSFC).
200211922/2
holes in Milky Way galaxy (Brown&Bethe1994; Timmesetal. 1996), however, only roughly two dozen dynamically-confirmed black hole X-ray binaries have been confirmed (Remillard&Mc- Clintock 2006), and 4U 1543-47 (4U 1543) is one of them. This transient source was first discovered by Uhuru satellite in 1971 (Matilsky et al. 1972). Then, it went into outburst again in 1983, 1992 and 2002, respectively (Kitamotoetal.1984; Harmonetal. 1992; Parketal.2004). It was a very long time after the first dis- covery that the compact primary was confirmed to be a black hole (Rhoades&Ruffini1974; Oroszetal.1998). As to its spin parameter, Shafeeetal. (2006) first reported its spin with the continuum-fitting method. They estimated its spin to be $0.8\pm0.1$. Then, and Morningstar&Miller (reported two spin measurements, $0.3\pm0.1$ and $0.43^{+0.22}_{-0.31}$, respectively, both constrained by combining the continuum-fitting and the X-ray reflection fitting methods. These three works utilized the previous mass of $9.4 \pm{1.0} M_{\odot}$ and distance of $7.5 \pm{1.0}$ kpc, which were reported in Parketal. (2004). Except that Milleretal. (2009) used the inclination angle of $32_{-4}^{+3}$ degrees constrained by their own fits to the iron line, the other two works used the inclination angle of $20.7 \pm{1.5}$ degrees which is equal to the binary orbital inclination angle. Recently, an updated set of dynamical parameters have been identified which have significant differences compared to earlier (J.Orosz, private communication). In this paper, we revisited 7 Rossi X-ray Timing Explorer (RXTE, Zhangetal.1993) data-sets of 4U 1543 to check the spin parameter via the X-ray reflection fitting method, i.e., carefully ex- ploring the reflection component. We made a joint-fit for all spectra in order to achieve a better signal-to-noise ratio. we have adopted the updated reflection-emission model, relxill (Garcíaetal.2014a; Dauser et al. 2014). The whole paper is organized as follows. In Section 2, we provide details of the data reduction and selection. In Section 3, we describe the analysis of spectra and the spin re- sult. In Section 4 and Section 5, we present our discussions and conclusions, respectively. ## DATA REDUCTION AND SELECTION We revisited data sets for 4U 1543 which were observed by RXTE during its 2002 outburst (Parketal.2004). There are 130 contin- uous pointed observations in total (the long exposures were split), collected by the Proportional Counter Array (PCA, Jahoda et al. 1996). We only focused our analysis on the best-calibrated propor- tional counter unit, namely PCU2, as in previous work (Parketal. 2004; Jahodaetal.2006; Shaposhnikovetal.2012). All layers of the PCU2 were combined. We neglected 49 observations whose count rate is smaller than 10 counts s $^{-1}$. The remaining observations are presented in the hardness-intensity diagram (HID, Figure 1). RXTE /PCA data of bright X-ray binaries are fundamentally limited not by counting statistics but by the systematic measure of calibration certainty in the detector. We apply a calibration correc- tion, pcacorr (Garcíaetal.2014b), which improves the instrumen- tal response to a quality of 0.1% precision. We include this 0.1% as a systematic error. A second correction, crabcorr (Steineretal. 2010), standardizes the PCA absolute flux calibration to the Toor & Seward (1974) values for the Crab. This latter tool improves not on the precision of the detector, but on the accuracy of our measurement. We firstly subtracted background and made deadtime correc- tion for the RXTE data. Next, the calibration tool pcacorr was applied. Then, a 0.1% systematic error was added to the spectra. Figure1. The evolutionary tracks in hardness-intensity diagram for obser- vations except those with count rate smaller than 10 counts s $^{-1}$. The vertical axis presents the count rate in energy band 3-45 keV. The horizontal axis presents the hardness ratio (HR) defined as the ratio of count rate between 5-8.6 keV and 8.6-18 keV. The 7 red open circles represent the data (two overlapping) we used to determine the spin of the black hole. Finally, we performed RXTE data analysis over the energy range between 2.8-45.0 keV using XSPEC 12.9.0g software package (Ar- naud1996). The quoted errors were given with a 90% confidence level ($\Delta \chi ^2=2.71$) if not specified. We selected 7 observations (MJD 52459-MJD 52463, defined as Spec. A-G) which show strong reflection components. In Table 1, we give the detailed information for these observations. In order to show the reflection features more clearly, we analysed the Spec. A-G between 2.8-45.0 keV, omitting 4.5-8.0 keV and 15.0-35.0 keV with the model crabcor*TBabs*(diskbb+powerlaw) in XSPEC. For the model crabcor, the normalization coefficient of $C=1.097$ and the slope difference of $\Delta\Gamma = 0.01$, are applied. For the model TBabs, which is used to account for the galactic absorption by the interstellar medium (ISM) along the line of sight, the Wilms et al. (2000) set of solar abundances and the Verner et al. (1996) photoelectric cross sections were specified accordingly. Since the effective low energy of RXTE is limited at 2.8 keV, the data cannot constrain the column density ($N_\mathrm{H}$) well. The column density 1 was fixed at $4.0\times10^{21}$ cm $^{-2}$ as in and Morningstar &Miller (2014). The fits to all 7 spectra are statistically unacceptable with $\chi^2_{\nu}$ = 31.32 (2129.78/68), 32.59 (2216.33/68), 23.26 (1581.93/68), 36.65 (2492.28/68), 63.75 (4335.57/68), 49.58 (3371.47/68) and 39.70 (2699.92/68), respectively. Data-to-model ratios are plotted in Fig- ure 2. The positive features in residuals are the broadened iron line and Compton hump characteristic of reflection emission. ## ANALYSIS AND RESULTS XMM-Newton 1 Milleretal. (2003) measured the column density for 4U 1543-47 to be (3.8 $\pm$ 0.2) $\times10^{21}$ cm $^{-2}$ by analysing the /EPIC-pn spectrum (0.3-10.0 keV). This value is consistent with the one we currently use in the fit. The new value of the column density has a negligible effect on our fitting results.
200211922/6
Figure3. Making joint-fit to Spec. A-G using model: crabcor*TBabs*(diskbb+relxill). Data-to-model ratios and contributions to the total $\chi^2$ are presented in left and right panels, respectively. Different colors represent different spectra. The model achieved a satisfactory fit with $\chi^2_{\nu}=390.40/452$. comes worse than when the inclination angle is free ($\Delta \chi^2 = 53.51$ for increasing 1 d.o.f). The best-fitting values are listed in Table 3. The temperature and the normalization of the thermal emission do not change significantly. The photon index parameters become smaller than the Newtonian value ($q=3$), which indicates the coro- nal model changes from a compact geometry ($q>3$ in Model 1) to extended. As for reflected emission, the emissivity index decreased. The spin pegged at 0.998 in this condition, possibly owning to the higher iron abundance of $7.72_{-1.62}^{+1.23}$, a more nonphysical value. The ionization state becomes much higher, and the reflection fraction becomes smaller. As an extension of Model 2, we define a new Model 3 in which we keep the inclination fixed at 21.0 $^{\circ}$ and also fix the spin parameter to the value found by Milleretal. (2009): $a_{*}$ =0.3. The best-fitting values are listed in Table 4. Compared to Model 2, its $\chi^2$ increased 12.99 for 1 d.o.f. Except the iron abundance was constrained at $4.50_{-0.21}^{+0.46}$, which is lower than that in Model 2, other parameters were not appreciably affected. We also plot the contribution to $\chi^2$ for Spec. A resulting from Model 1-3 in Figure 9. The most pronounced changes in comparing the model differences are residuals around the iron line region ($\sim$ 5.0-8.0 keV). ## CONCLUSIONS We have measured the spin of 4U 1543 via modeling its reflected components in 7 SPL state observations carefully. The spectra con- sist of 4 different components: the galactic absorption, thermal emis- sion from the accretion disc, power-law emission and reflected emis- sion. We did joint-fit to all the spectra to improve the signal-to-noise ratio of X-ray reflected component. We used the reflection model, relxill, to fit the data. We find a super-solar iron abundance for the disc. At the same time, the disc is highly ionized. The model with free inclination angle and spin (Model 1) de- scribes the spectra best. The inclination angle of the inner accretion disc is constrained to be $36.3_{-3.4}^{+5.3}$ degrees at 90 statistical confi- dence. When the inclination angle is fixed at the orbital inclination value of 21.0 $^{\circ}$ in Model 2 or Model 3, the statistic becomes sig- nificantly worse, and the spin is larger than 0.83. The best-fitting inclination differs from that of the orbital plane by more than 10 degrees. This may be owed to the systematic limitations of current models which underestimate the density of disc. Our results indicate a moderate rotation rate for the black hole in 4U 1543. The spin parameter is established to be $0.67_{-0.08}^{+0.15}$ at 90% statistical confidence. At the 99% statistical confidence level, we exclude spins below $a_{*}$ < 0.5 (which also excludes any retrograde geometries).
README.md exists but content is empty. Use the Edit dataset card button to edit it.
Downloads last month
0
Edit dataset card