arXiv:2107.01242v1 [math.OA] 2 Jul 2021 euneo ievle ( eigenvalues of sequence a eetdacrigt t agbac utpiiy y[ By multiplicity. (algebraic) its to according repeated nerli ocmuaiegoer.Aprilase oti que this to answer partial [ Sukochev-Zanin A of preprint geometry. noncommutative in integral limits? extended using without hc a eda ua nvriyi hnhi hn,fo ac 29- March from China, A Question Shanghai, in question: University following the Fudan asked at Connes held was which where en h Cintegral NC the define positive ektaecasoeaos ealif re Recall additivity operators. asymptotic class interesting trace an weak there establish authors [ the Lord-Sukochev-Zanin of book question. intermediate Connes’ some to at answer limits satisfactory extended a using provide on not relies still approach the [ cla Connes trace Following weak limits. Tr the extended traces on with A trace associated the are positive by traces a given These m be are quantum should choices of setup Natural language this theoretic in operator integral the into calculus itesimal ∈ nohrwrs onsi tesn h edfraprl spectra purely a for need the stressing is Connes words, other In uigtecneec Nnomttv emty tt fthe of state geometry: “Noncommutative conference the During h unie aclso ons[ Connes of calculus quantized The Date u by question Connes’ answer can we that observe we paper this In L 1 λ uy6 2021. 6, July : then , , 0 ∞ ocmuaiegoer n eilsia analysis. princip semiclassical Birman-Schwinger and the geometry of noncommutative trace W consequence Connes’ easy semiclassical an of and is form This integral stronger Connes’ a between yields relationship law Weyl’s pseudodiff This order negative fold. for Connes law the of Weyl’s perturbation relationship Birman-Solomyak’s Birman-Solomyak’s the Al and clarify operators of further integrcompact question We Lebesgue’s a with Connes. answers traces Alain Dixmier This of compatibility integral. the presen Connes’ we First, of geometry. struction noncommutative Connes’s of work Abstract. ( A smaual hntevleo Tr of value the when measurable is ) ≥ (Connes) λ  1 A ( A hspprda ihsm usin eadn h oino i of notion the regarding questions some with deal paper This ONS NERTO N ELSLAWS WEYL’S AND INTEGRATION CONNES’ smaual and measurable is ) · · · ≥ . R − si osbet hwteeitneo ii o l measura all for limit a of existence the show to possible it Is A λ ob hsvle tflosfo hscntuto htif that construction this from follows It value. this be to j r h ievle of eigenvalues the are 61 ( A .Hwvr htpprdaswt pca ls foeaosand operators of class special a with deals paper that However, ]. 34 )) j .Atog h anfcso hsbo so iglrtraces, singular on is book this of focus main the Although ]. ≥ 0 A uhthat such 18 Z − 1. RAPHA isa rnltn h antoso h lsia infin- classical the of tools main the translating at aims ] sacmatoeao,issetu a eognzdas organized be can spectrum its operator, compact a is A Introduction = ω ω L fDxir[ Dixmier of ( LPONGE EL  ¨ A | 1 λ sidpneto h xeddlmt ethen We limit. extended the of independent is ) ⇐⇒ 0 A ( A one ihmultiplicity. with counted ) N | ≥ | 34 lim →∞ em ..]if 5.7.5] Lemma , l hsasesaohrqeto of question another answers This al. y’ a o Schr¨odinger operators. for law eyl’s λ 23 e etu e etln between link neat a get thus We le. r.W logv sf ro”of proof” ”soft a give also We ory. rniloeaoso lsdmani- closed on operators erential 1 nerto ihWy’ asfor laws Weyl’s with integration ’ log ueyseta hoei con- theoretic spectral purely a t ( hoe.Fnly eepanthe explain we Finally, theorem. seas [ also (see ] i ons eas elwith deal also We Connes. ain A 1 N ) · · · ≥ | j

(1.1) λj (A + B)= λj (A)+ λj (B) + O(1). jX≤N jX≤N jX≤N This result is related to an eigenvalue characterization of the commutator space Com(L1,∞). It is also used in [34] to show that an operator A L1,∞ is measurable if and only it is Tauberian, in the sense that ∈ 1 (1.2) lim λj (A) exists. N→∞ log N j

Trω A(x) dµ(x)= Trω A(x)dµ(x) . ZΩ  ZΩ  In particular, if (Ω,µ) is a finite measure  space, then the above result holds for any (essentially) bounded measurable map A : Ω L1,∞. Furthermore, for such maps and in the special case of Dixmier traces associated with→ medial limits, this result is an immediate consequence of the fundamental property of medial limits alluded above (see Section 3). This confirms Connes’ suggestion. As it turns out, there are numerous positive traces on L1,∞ that are not Dixmier traces. Therefore, it is natural to consider a notion of measurability with respect to all positive traces on L1,∞ (see, e.g., [29, 34, 55]). We shall call such operators strongly measurable. For sake of 2 completeness we overview their main properties. These operators actually form a natural domain for the NC integral, in the sense that its restriction to strongly measurable still satisfies the NC integral’s Ansatz (see Proposition 4.5). We also establish the spectral invariance of the strong measurability property (see Proposition 4.9). Strong measurability naturally appear in the context of Connes’ trace theorem [17, 29]. Suppose that (M n,g) is a closed Riemannian manifold and E is a Hermitian vector bundle over M. Recall that Connes’ trace theorem asserts that if P : C∞(M, E) C∞(M, E) is a pseudodifferential operator (ΨDO) of order n, then P is strongly measurable,→ and we have − 1 (1.4) P = trE σ(P )(x, ξ) dxdξ, − n ∗ Z ZS M where σ(P ) is the principal symbol of P and S∗M is the cosphere bundle equipped with its Liouville measure dxdξ. The r.h.s. is the noncommutative residue trace of P in the sense of Guillemin [28] −n/2 ∞ and Wodzicki [65]. Applying the above result to P = f∆g , where f C (M) and ∆g is the Laplace-Beltrami operator on functions, gives Connes’ integration formula,∈

n − 2 1 −n n−1 (1.5) f∆g = c f(x) g(x)dx, c := (2π) S . − n n n Z ZM p This shows that the NC integral recaptures the Riemannian measure g(x)dx . In the special case f = 1 Connes’ integration formula (1.5) is an immediate consequence of the p Weyl’s law for the Laplacian ∆g. Strong measurability does not imply Weyl’s law. Therefore, we are lead to the following question: Question C. What is the precise relationship between Weyl’s law and measurability? This question is closely related to the work of Birman-Solomyak [7, 10, 11, 12] on Weyl’s laws for compact operators in the 70s. This work was partly motivated by the semiclassical analysis of Schr¨odinger operators, since by the Birman-Schwinger principle Weyl’s laws for compact operators yield semiclassical Weyl’s laws for Schr¨odinger operators. In particular, Birman-Solomyak [7] set- up a full perturbation theory. In [10, 11, 12] they further showed that if P is a selfadjoint ΨDO of order m< 0, and we set p = nm−1, then we have the following Weyl’s law, − 1 1 1 p p ± −n p (1.6) lim j λj (P )= (2π) trE σ(P )(x, ξ)± dxdξ . j→∞ n ∗  ZS M  where λ±(P ) λ±(P ) ... are the positive and negative eigenvalues of P . We also have ± 0 ≥ ± 1 ≥ a similar results for the singular values of P (i.e., the eigenvalues of P = √P ∗P ) without any selfadjointness assumption. Namely, | | 1 p 1 1 −n p (1.7) lim j p µj (P )= (2π) trE σ(P )(x, ξ) dxdξ , j→∞ n ∗ | |  ZS M  where µ (P ) µ (P ) are the singular values of P.  0 ≥ 1 ≥ · · · As it turns out (see Proposition 5.12), any selfadjoint operator A L1,∞ satisfying a Weyl’s law of the form (1.6) for p = 1 is strongly measurable, and we have ∈

(1.8) A = lim jλ+(A) lim jλ−(A). − j→∞ j − j→∞ j Z There is a similar result for the absolute value A in terms of the singular values of A (cf. Corol- lary 5.13). This answers Question C. This also| provides| a further spectral theoretic description of the NC integral for operators satisfying Weyl’s laws. Incidentally, this shows that the Weyl’s laws (1.6)–(1.7) of Birman-Solomyak provide us with a stronger form of Connes’ trace theo- rem (1.4). For instance, if P is any ΨDO of order n, then its absolute value P is strongly measurable, even though it need not be a ΨDO. − | | The original proof of the Weyl’s laws (1.6)–(1.7) by Birman-Solomyak [7, 10, 11, 12] is arguably a beautiful piece of hard analysis. Unfortunately, the main key technical details are exposed in a somewhat compressed manner in the Russian article [11], the translation of which remains unavailable. 3 Question D. Is there a “soft proof” of Birman-Solomyak’s Weyl’s laws (1.6)–(1.7)? We provide such a proof in Section 6. The approach uses the relationship between zeta functions and the noncommutative residue to get the Weyl’s laws (1.6)–(1.7) for inverses of elliptic operators. The perturbation theory of Birman-Solomyak [7, §4] and the BKS inequality [6] then allow us to get the Weyl’s laws in the general case. We refer to Section 6 for the full details. and semiclassical analysis are usually considered to be different sub-fields of quantum theory. As mentioned above the Birman-Schwinger principle provides a bridge between Weyl’s laws for compact operators and semiclassical Weyl’s laws for Schr¨odinger operators. As we have related the former to Connes’ integration, we are lead to the following question: Question E. What is the precise relationship between Connes’ integral and semiclassical Weyl’s laws for Schr¨odinger operators? It’s clear that an answer is provided by the Birman-Schwinger principle and the integration for- mula (1.8). The question is more like to determine the level of generality at which the relationship holds and explain how simple this relation is (compare [39]). To wit let H be a bounded from below operator with non-negative spectrum. We assume that 0 is an isolated eigenvalue with finite multiplicity. Let V be a selfadjoint H-form compact perturbation. The operator H + V then makes sense as a form sum and the negative part of its spectrum is discrete. Let N −(H + V ) be the number of its negative eigenvalues counted with multiplicity (i.e., the number of bound states in physics’ jargon). The abstract Birman-Schwinger principle [13] (see also [36]) relates N −(H +V ) to the counting function of the Birman-Schwinger operator H−1/2VH−1/2. Combining this with the integration formula (1.8) shows that if V 0 and H−1/2VH−1/2 satisfies a Weyl’s law of the form (1.6) for p = 1, then (see Proposition 7.3≥) we have

2 2 1 1 lim h N − h H V = H− 2 VH− 2 . h→0+ − − Z This answers Question E and gives a neat link between Connes’ noncommutative geometry and the semiclassical analysis of Schr¨odinger operators. We also illustrate this result in terms of the recent Weyl’s laws for LlogL-Orlicz potentials on closed manifolds by Rozenblum [49] and Sukochev- Zanin [61] (see also [45, 50]). This leads us to a semiclassical interpretation of Connes’ integration formula (1.5) (see Section 7). The remainder of this paper is organized as follows. In Section 2, we deal with Question A and give a purely spectral theoretic construction of Connes’ integral. In Section 3, we deal with Question B. In Section 4, we describe the main properties of strongly measurable operators. In Section 5, we deal with Question C by relating Connes’ integration to the Weyl’s laws for compact operators. In Section 6, we give a “soft proof” of Birman-Solomyak’s Weyl’s laws (1.6)–(1.7); this deals with Question D. In Section 7, we deal with Question E by explaining the link between Connes’ integral and semiclassical Weyl’s laws. Finally, in Appendix A, we gather a few results on Hilbert spaces embeddings that are needed in Section 2.

2. Quantized Calculus and NC Integral In this section, we present a purely spectral theoretic construction of Connes’ integral. After a brief review of weak Schatten classes and Connes’ quantized calculus, we give two constructions of the NC integral. The first one is given in terms of Dixmier traces and uses extended limits. The other construction is in terms of Tauberian operators. These two constructions are shown to give exactly the same notion of NC integral. This will answer Question A. 2.1. Weak Schatten Classes. We briefly review the main definitions and properties regarding Schatten and weak Schatten classes (see, e.g., [59, 26] for further details). Throughout this paper we let H be a (separable) with inner product . The algebra of bounded linear operators on H is denoted L (H ). The operator norm is denotedh·|·i . We also denote by K the (closed) ideal of compact operators on H . Given any operator T k·kK ∈ 4 we let (µj (T ))j≥0 be its sequence of singular values, i.e., µj (T ) is the (j +1)-th eigenvalue counted with multiplicity of the absolute value T = √T ∗T . The min-max principle states that | | (2.1) µ (T ) = min T ⊥ ; dim E = j . j k |E k We record the following properties of singular values (see, e.g., [26 , 59]), (2.2) µ (T )= µ (T ∗)= µ ( T ), j j j | | (2.3) µ (S + T ) µ (S)+ µ (T ), j+k ≤ j k (2.4) µ (AT B) A µ (T ) B , A,B L (H ),. j ≤k k j k k ∈ The inequality (2.3) is known as Ky Fan’s inequality. p For p (0, ) the Schatten class Lp consist of operators T K such that T is trace-class. It is equipped∈ ∞ with the quasi-norm, ∈ | |

1 1 p T := Tr T p p = µ (T )p ,T L . k kp | | j ∈ p  j≥0   X We obtain a quasi-Banach ideal. For p 1 the L -quasi-norm is actually a norm, and so in this ≥ p case Lp is a Banach ideal. In any case, the finite-rank operators on H form a dense subspace of Lp. For p (0, ), the weak Schatten class L is defined by ∈ ∞ p,∞ − 1 L := T K ; µ (T )=O j p . p,∞ ∈ j This is a two-sided ideal. We equip itn with the quasi-norm, o

1 (2.5) T p,∞ := sup (j + 1) p µj (T ),T Lp,∞. k k j≥0 ∈ For p> 1, the quasi-norm is equivalent to the norm, k·kp,∞ 1 ′ −1+ p L T p,∞ := sup N µj (T ),T p,∞. k k N≥1 ∈ j

1 (2.6) S + T 2 p ( S + T ) , S,T L p,∞. k kp,∞ ≤ k kp,∞ k kp,∞ ∈ In addition, we denote by (Lp,∞)0 the closure in Lp,∞ of the finite-rank operators. We have

− 1 L = T K ; µ (T )=o j p . p,∞ 0 ∈ j We note the continuous inclusions,  n o L L L L p ( p,∞ 0 ( p,∞ ( q, 0 < p < q.

In the following, we will also denote the Schatten and weak Schatten classes by Lp(H ) and Lp,∞(H ) whenever there is a need to specify the Hilbert space.

2.2. Quantized calculus. The main goal of the quantized calculus of Connes [18] is to translate into the Hilbert space formalism of quantum mechanics the main tools of the classical infinitesimal calculus. Classical Quantum Complex variable Operator on H Real variable Selfadjoint operator on H Infinitesimal variable on H Infinitesimal of order α> 0 Compact operator T such that −α µj (T )=O(j ) 5 The first two lines arise from quantum mechanics. Intuitively speaking, an infinitesimal is meant to be smaller than any real number. For a the condition T < ǫ for all ǫ > 0 gives T = 0. This condition can be relaxed into the following: For every kǫ >k 0 there is a finite-dimensional subspace E of H such that T|E⊥ < ǫ. This is equivalent to T being a compact operator. k k The order of compactness of a compact operator is given by the order of decay of its singular values. Namely, an infinitesimal operator of order α > 0 is any compact operator such that −α −1 µj (T ) = O(j ). Thus, if we set p = α , then T is infinitesimal operator of order α > 0 iff T Lp,∞. ∈The next line of the dictionary is the NC analogue of the integral. As an Ansatz the NC integral should be a linear functional satisfying at least the following conditions: (1) It is defined on a suitable class of infinitesimal operators of order 1. (2) It vanishes on infinitesimal operators of order > 1. (3) It takes non-negative values on positive operators. (4) It is invariant under Hilbert space isomorphisms. As mentioned above, the infinitesimal operators of order 1 are the operators in the weak L1,∞. The condition (3) means that the functional should be positive. The condition (4) forces the functional to be a trace, in the sense it is annihilated by the commutator subspace, Com L := Span [A, T ]; A L (H ),T L . 1,∞ { ∈ ∈ 1,∞} More precisely, it would be convenient to adopt the following definition of a trace.

Definition 2.1. If E is a subspace of L1,∞ containing Com(L1,∞) and F is another vector space, then we say that a linear map ϕ : E F is a trace if it is annihilated by Com(L ). → 1,∞ To sum up the NC integral should be a positive trace − : M C, where M is a suitable → subspace of L1,∞ containing the commutator subspace Com(L1,∞) and infinitesimal operators of order > 1. R

2.3. Eigenvalue sequences and commutators in L1,∞. If A is a compact operator on H , then its spectrum can be arranged as a sequence (λj (A))j≥0 converging to 0 such that λ (A) λ (A) λ (A) 0, | 0 | ≥ | 1 | ≥ · · · ≥ | j |≥···≥ where each eigenvalue is repeating according to its algebraic multiplicity, i.e., the dimension of the root space E (A) := ker(A λ)ℓ. If λ = 0, the algebraic multiplicity is always finite (see, λ ℓ≥0 − 6 e.g., [26]). It agrees with the geometric multiplicity whenever A is normal. S A sequence as above is called an eigenvalue sequence for A. Such a sequence is not unique. If A 0, then the eigenvalue sequence is unique and agrees with its singular value sequence ≥ (µj (T ))j≥0. In general, an eigenvalue sequence need not be unique. In what follows by (λj (A))j≥0 we shall always denote an eigenvalue sequence in the sense above. We record the Ky Fan’s inequalities (see, e.g., [26, 59]).

(2.7) λ (A) λ (A) µ (A) N 1. j ≤ | j |≤ j ∀ ≥ j

Lemma 2.2 ([34, Lemma 5.7.5]). If A and B are operators in L1,∞, then

(2.8) λj (A + B)= λj (A)+ λj (B) + O(1). j

′ (2.9) λj (A)= λj (A) + O(1). j

λj (A) = O(1). j

λ (A)= λ (U ∗BU) λ (B) + O(1) = O(1). j j − j j

We have a converse to Corollary 2.6. More precisely, we have the following result.

Proposition 2.7 ([24, 34]). If A and B are operators in L1,∞, then

(2.10) A B Com L λ (A)= λ (B) + O(1). − ∈ 1,∞ ⇐⇒ j j j 1.

Proof. Let A L1,∞, and let (λj (A))j≥0 be an eigenvalue sequence. If c C, then (cλj (A))j≥0 is an eigenvalue∈ sequence for cA, and hence ∈ 1 τ(cA) = class of cλj (A) = cτ(A). log N N≥1  j

Therefore, we see that τ is a continuous linear map. It is immediate that if A has non-negative eigenvalues, then τ(A) is a positive element of ℓ∞/ℓ0. Thus, τ is a positive linear map. Furthermore, it follows from Corollary 2.6 that if A Com(L ), then (log N)−1 λ (A) is o(1), and hence τ(A) = 0. Thus, τ is a trace. ∈ 1,∞ j

(2.13) lim aN = L a L ℓ0 limω a = L ω . N→∞ ⇐⇒ − ∈ ⇐⇒ ∀ 8   Given any extended limit lim we define Tr : L C by ω ω 1,∞ → 1 Tr (A) = lim λ (A), A L . ω ω log N j ∈ 1,∞ j 1. Definition 2.11. The trace Tr : L C is called the Dixmier trace associated with the ω 1,∞ → extended limit limω. Every Dixmier trace satisfies the Ansatz for the NC integral. However, if A L , the value ∈ 1,∞ of Trω(A) may depend on the choice the extended limit. To remedy this we proceed as follows.

Definition 2.12 (Connes [18]). An operator A L1,∞ is called measurable if the value of Trω(A) is independent of the choice of the extended limit.∈ For such an operator, its NC integral is defined by 1 A = lim λ (A), − ω log N j Z j 1 are measurable. (2) The NC integral − : M C is a positive continuous trace on M . It is annihilated by operators in (L ) , including→ infinitesimal operators of order > 1. 1,∞R 0 Proof. By definition, M = A L ; Tr (A) = Tr ′ (A) , ∈ 1,∞ ω ω ω,ω′ ′ \  where ω and ω range over all states on ℓ∞/ℓ0. As the Dixmier traces Trω are continuous linear maps, it follows that M is a closed subspace of L1,∞. By definition the NC integral − agrees with any Dixmier trace Trω, and so this is a continuous positive linear functional by Proposition 2.10. Moreover, as the union Com(L ) (L ) is R 1,∞ ∪ 1,∞ 0 annihilated by every Dixmier trace, it is contained in M and is annihilated by −. In particular, the NC integral − is a trace on M . The proof is complete.  R 2.5. A noncommutativeR integral in terms of Tauberian operators. We shall now present an alternative approach to the NC integral. The approach is to work with Tauberian operators (see definition below). This approach is inspired by the characterization of measurable operators in terms of Tauberian operators in [34]. We will show in the next subsection that this approach is equivalent to the previous approach in terms of Dixmier approach. As the 2nd approach involves spectral data only, the equivalence between the two approaches will answer Question A. L ′ If A 1,∞, then it follows from (2.9) that if (λj (A))j≥0 and (λj (A))j≥0 are two eigenvalue sequences∈ for A, then −1 ′ −1 (log N) λj (A) = (log N) λj (A) + o(1). j

(1) T is a subspace of L1,∞ containing Com(L1,∞) and (L1,∞)0. ′ (2) The map − : T C is a continuous positive linear trace on T . It is annihilated by operators in (L →) , including infinitesimals of order > 1. R 1,∞ 0 ′ Proof. It is immediate that if an operator A T is positive, then − A 0. Moreover, if ∈ ≥ A T and c C, then (cλj (A))j≥0 is an eigenvalue sequence of cA, and hence cA T with ′ ∈ ′ ∈ R ∈ − (cA)= c− (A). If A, B T , then (2.8) implies that ∈ 1 1 1 R R λ (A + B)= λ (A)+ λ (B) + o(1). log N j log N j log N j j

Theorem 2.17. An operator A L1,∞ is measurable if and only if it is Tauberian. Moreover, in this case we have ∈ 1 (2.14) A = lim λj (A). − N→∞ log N Z j

Proposition 2.19. Let A L1,∞. Then A is measurable if and only if its real part A and its imaginary part A are both∈ measurable. Moreover, in this case we have ℜ ℑ A = A, A = A. ℜ − − ℜ ℑ − − ℑ Z  Z Z  Z Proof. It follows from Remark 2.5 that 1 1 λ (A) = λ ( A) + o(1), log N ℜ j log N j ℜ j

Proposition 2.20. Let A = A∗ L . Then A is measurable if and only if ∈ 1,∞ 1 + − lim (λj (A) λj (A)) exists. N→∞ log N − j

1 + − A = lim (λj (A) λj (A)). − N→∞ log N − Z j

Let H ′ be another Hilbert space. Theorem 2.17 implies the following spectral invariance result.

′ ′ Proposition 2.21. Let A L1,∞(H ) and A L1,∞(H ) have the same non-zero eigenvalues with same multiplicities. Then∈ A is measurable∈ if and only if A′ is measurable. Moreover, in this case − A = − A′. SupposeR R now that ι : H ′ H is continuous linear embedding, i.e., it is a linear map which is one-to-one and has closed range.→ For instance, any isometric linear map is such an embedding. Denote by H1 the range of ι. By assumption this is a closed subspace of H and ι gives rise to −1 ′ a continuous linear isomorphism ι : H H1 with inverse ι : H1 H . As explained in Appendix A we have a pushforward map →ι : L (H ′) L (H ) given by→ ∗ → (2.15) ι A = ι A ι−1 π, A L (H ′), ∗ ◦ ◦ ◦ ∈ 11 where π : H H is the orthogonal projection onto H1. In particular, if ι is invertible, then −1 → ι∗A = ιAι . We also know from Proposition A.2 that ι∗ induces a continuous linear embedding, ι : L (H ′) L (H ). ∗ 1,∞ −→ 1,∞ ′ Moreover, by Proposition A.1 if A L1,∞(H ), then A and ι∗A have the same non-zero eigenval- ues with same multiplicities. Combining∈ this with Proposition 2.21 we then arrive at the following statement.

′ Corollary 2.22. Let A L1,∞(H ). Then ι∗A is measurable if and only A is measurable. Moreover, in this case we∈ have ι A = A. − ∗ − Z Z Denote by M (H ) (resp., M (H ′)) the space of measurable operators on H (resp., H ′). Specializing Corollary 2.22 to the case where ι is an isomorphism yields the following invariance result. Corollary 2.23. Assume ι : H H ′ is a continuous linear isomorphism. Then ιM (H )ι−1 = M (H ′), and we have → ιAι−1 = A A M (H ). − − ∀ ∈ Z Z 3. Connes’ Integration and Lesbegue Integration In this section, we look at the compatibility of Connes’ integral with Lebesgue’s integration. This will answer Question B.

3.1. Compatibility of Dixmier traces with Lebesgue’s integration. To address the com- patibility of Dixmier traces with Lebesgue’s integration, the main technical hurdle is the lack of convexity of the weak trace class L1,∞. Indeed, L1,∞ is a quasi-Banach ideal, but this is not a Banach space or even a locally convex space. Thus, Bochner integration, or even Gel’fand-Pettis integration, of maps with values in L1,∞ do not make sense. We can remedy this by passing to the closure L 1,∞ in the Dixmier-Macaev ideal L(1,∞). Recall that L := A K ; µ (T ) = O(log N) . (1,∞) ∈ j j 0, such that, for all A L , we have ∈ 1,∞ 1 1 τ(A) sup λ (A) µ (A) C A , k k≤ log N j ≤ log N j ≤ k k(1,∞) N≥1 j

Thus, the linear map τ is continuous with respect to the Dixmier-Macaev norm, and hence it uniquely extends to a continuous linear map τ : L 1,∞ ℓ∞/ℓ0. This map is positive and is a trace. The proof is complete. →  12 Given any state ω on ℓ /ℓ , we define the map Tr : L C by ∞ 0 ω 1,∞ → Tr (A)= ω τ(A), A L . ω ◦ ∈ 1,∞ Equivalently, Trω is the unique continuous extension to L 1,∞ of the Dixmier trace Trω. In what follows we let (Ω,µ) be a measure space. L Proposition 3.2. Let A :Ω 1,∞ be a measurable map s.t. Ω A(x) (1,∞)dµ(x) < . Then, for every extended limit lim ,→ the function Ω x Tr [A(x)] is integrable,k k and we have∞ ω ∋ → ω R

(3.2) Trω A(x) dµ(x)= Trω A(x)dµ(x) . ZΩ  ZΩ    In particular, if A(x)dµ(x) L , then Ω ∈ 1,∞ R Trω A(x) dµ(x) = Trω A(x)dµ(x) . ZΩ  ZΩ    Proof. Let lim be an extended limit. The continuity of Tr and the fact that A : Ω L is ω ω → 1,∞ Bochner-integrable as an L 1,∞-valued map ensure us that the function Trω[A(x)] = Trω[A(x)] is integrable, and we have

Trω A(x) dµ(x)= Trω A(x) dµ(x)= Trω A(x)dµ(x) . ZΩ ZΩ  ZΩ  The proof is complete.    

Corollary 3.3. Let A : Ω L be a measurable map so that A(x) dµ(x) < . → 1,∞ Ω k k(1,∞) ∞ Assume that A(x) is a measurable operator a.e., and A(x)dµ(x) L . Then A(x)dµ(x) Ω R 1,∞ Ω is a measurable operator, and we have ∈ R R A(x) dµ(x)= A(x)dµ(x) . − − ZΩ Z  Z  ZΩ  3.2. Dixmier traces associated with medial limits. As pointed out by Connes [19] another route to look at the compatibility of Connes’ integration with Lebesgue integration is to use medial limits. These limits were introduced by Mokodoski [41]. Namely, by using the continuum hypothesis he proved the following result. Lemma 3.4 (Mokodoski [41]). There exists a state ω : ℓ /ℓ C which is universally measurable ∞ 0 → and such that, for any complete finite measure µ on ℓ∞/ℓ0, we have

(3.3) ω adµ(a) = ω(a)dµ(a).  Z  Z Let ωmed be a state as in the above lemma. The corresponding extended limit is called a medial limit and is denoted by lim med. The fundamental property (3.3) implies the following striking feature of medial limits.

Proposition 3.5 ([41]). Given a complete finite measure space (Ω,µ), let (fℓ)ℓ≥1 a bounded family in L∞(Ω,µ).

(i) The function Ω x lim med fℓ(x) is measurable. (ii) We have ∋ →

lim med fℓ(x) dµ(x) = lim med fℓ(x)dµ(x). Zω ZΩ  Proposition 3.6 ([41]). Given a complete finite measure space (Ω,µ), let (fℓ)ℓ≥1 a bounded family in L∞(Ω,µ). Then Ω x lim med f (x) is a bounded measurable function such that ∋ → ℓ

lim med fℓ(x) dµ(x) = lim med fℓ(x)dµ(x). Zω ZΩ  13 In other words, we don’t have to worry about the integrability of lim med fℓ(x). We may freely swap the integral sign and the medial limit. Alternatively, if a : Ω ℓ∞/ℓ0 is any bounded measurable map, then →

(3.4) ωmed a(x) dµ(x)= ωmed a(x)dµ(x) . ZΩ  ZΩ  Denote by Tr the Dixmier trace  associated with the extended limit lim med. Let A :Ω ωmed → L1,∞ be a bounded measurable map. Applying (3.4) to a(x)= τ[A(x)] gives

ω τ A(x) dµ(x)= ω τ A(x)]dµ(x) . med ◦ med Z  ZΩ  Note that ω τ[A(x)] = Tr [A(x)], and  med ◦ ωmed τ A(x)]dµ(x)= τ A(x)]dµ(x)= τ A(x)dµ(x) . ZΩ ZΩ  ZΩ  Thus,  

Tr A(x)]dµ(x)= ω τ A(x)dµ(x) = Tr A(x)dµ(x) . ωmed med ◦ ωmed Z  ZΩ   ZΩ  Therefore, in the special case of medial limits, we recover the formula (3.2) as an immediate consequence of the fundamental property (3.3) of those extended limits.

4. Strongly Measurable Operators In this section, we look at a stronger notion of measurability and show that we still have a sensible notion of integral on operators that satisfies this notion of measurability.

4.1. Strong measurability. In what follows we denote by T0 any positive operator in L1,∞ such that λ (T ) = (j + 1)−1 for all j 0. Note that any two such operators are unitary equivalent, j 0 ≥ and hence agree up to an element of Com(L1,∞). Recall that a trace ϕ : L C is called normalized if ϕ(T ) = 1. All the Dixmier traces are 1,∞ → 0 normalized traces. However, there are positive normalized traces on L1,∞ that are not Dixmier traces and don’t have a continuous extension to the Dixmier-Macacev ideal L(1,∞) (see, e.g., [55, Theorem 4.7]). Therefore, it stands for reason to consider a stronger notion of measurability (see, e.g., [29, 34, 55]). Definition 4.1. An operator T L is called strongly measurable when there is L C such ∈ 1,∞ ∈ that ϕ(T ) = L for every positive normalized trace ϕ on L1,∞. We denote by Ms the class of strongly measurable operators. Remark 4.2. The class of strongly measurable operators is strictly contained in the space of measurable operators (see [55, Theorem 7.4]).

Lemma 4.3. The space of continuous traces on L1,∞ is spanned by normalized positive traces. In fact, any continuous traces is a linear combination of 4 normalized positive traces.

Proof. Every positive trace on L1,∞ is continuous (see, e.g., [44, Proposition 2.2]). Conversely, every continuous trace on L1,∞ is a linear combination of 4 positive traces (see [20, Corollary 2.2]). To complete the proof it is enough to show that every positive trace is a scalar multiple of a normalized positive trace. Let ϕ be a non-zero positive trace. As L1,∞ is spanned by its positive cone, there is a positive operator A L1,∞ such that ϕ(A) > 0. Let (ξj )j≥0 be an orthonormal basis of H such that ∈ −1 Aξj = µj (T )ξj for all j 0. Let T0 be the operator on H such that T0ξj = (j + 1) ξj . As −≥1 µj (A) A 1,∞(j + 1) we see that A A 1,∞T0. The positivity of ϕ then implies that 0 < ϕ(A≤) k kT ϕ(T ). Thus, ϕ(T ) > 0,≤ and k sok ϕ ˜ := ϕ(T )−1ϕ is a normalized positive trace. ≤k k1,∞ 0 0 0 As ϕ = ϕ(T0)˜ϕ we see that every positive trace is a scalar multiple of a normalized positive trace. The proof is complete.  14 Lemma 4.3 implies the following characterization of strongly measurable operators in terms of continuous traces. Lemma 4.4. Let A L . The following are equivalent: ∈ 1,∞ (i) A is strongly measurable and − A = L. (ii) ϕ(A)= ϕ(T )L for every continuous trace on L . 0 R 1,∞ This implies the following properties. Proposition 4.5. The following holds.

(1) Ms is a closed subspace of M containing Com(L1,∞) and (L1,∞)0. In particular, every infinitesimal operator of order > 1 is strongly measurable. (2) The space Ms does not depend on the inner product of H . (3) Let A L be such that ∈ 1,∞

(4.1) λj (A)= L log N + O(1). j

Proof. It is immediate that Ms is a subspaceR of M . Moreover, as the condition (ii) of Lemma 4.4 depends only on the topological vector space structure of L1,∞, we see that Ms does not depend on the inner product of H . Let (Aℓ)ℓ≥0 be sequence in Ms converging to A in L1,∞. Set αℓ = − Aℓ. As M is a closed subspace of L and − is a continuous linear form, we see that α α as ℓ . Let ϕ be a 1,∞ ℓ → R → ∞ positive normalized trace on L . We have ϕ(A )= − A = α for all ℓ 0. As ϕ is a continuous R 1,∞ ℓ ℓ ℓ trace, we have ≥ R ϕ(A) = lim ϕ(Aℓ) = lim αℓ = α. ℓ→∞ ℓ→∞ Thus ϕ(A)= α for every positive normalized trace, and so A is strongly measurable. This shows that Ms is a closed subspace of M . Furthermore, if A Com(L1,∞) and ϕ is a continuous trace on L1,∞, then ϕ(A) = 0, and so by using Proposition 4.4∈ we deduce that A is strongly measurable. It follows from Proposition 2.7 that the ideal R of finite rank operators is contained in Com(L1,∞). As Ms is closed, we deduce that it contains the closure of R in L1,∞, i.e., the ideal (L1,∞)0. Finally, let A L1,∞ satisfy (4.1). In addition, let T0 be any positive operator in L1,∞ such that µ (T ) = (j + 1)−∈1. Then (4.1) is the r.h.s. of (2.10) for B = LT , and so A LT Com(L ). j 0 0 − 0 ∈ 1,∞ In particular, given any normalized positive trace ϕ on L1,∞ we have ϕ(A)= Lϕ(T0)= L. That is, A M and − A = L. The proof is complete.  ∈ s

Remark 4.6. LetR us called strongly Tauberian any operator A L1,∞ satisfying (4.1). It same way as in the proof of Proposition 2.16, it follows from Lemma∈2.2 that the strongly Tauberian operators form a subspace of Ms. This is a proper subspace. For instance, every operator in Com(L ) Com(L ) is strongly measurable, but is not strongly Tauberian. 1,∞ \ 1,∞ Remark 4.7. As Ms is a subspace of M containing Com(L1,∞), we see that the NC integral − induces a positive continuous trace on Ms which is annihilated by operators in (L1,∞)0. This restriction too satisfies the Ansatz for the NC integral. R Remark 4.8. We refer to [55, Proposition 7.2] for a characterization of strongly measurable oper- ators in terms of eigenvalue sequences.

′ In what follows we denote by H another Hilbert space. We also denote by Ms(H ) (resp., ′ ′ Ms(H )) the space of strongly measurable operators on H (resp., H ). We have the following invariance property of strong measurable operators.

′ Proposition 4.9. Let A L1,∞(H ) and B L1,∞(H ) have same non-zero eigenvalues with same multiplicities. Then ∈A is strongly measurable∈ if and only if B is strongly measurable. 15 Proof. By assumption any eigenvalue sequence for A is an eigenvalue sequence for B, and vice versa. If H ′ = H , then combining this with Corollary 2.6 shows that A B is an operator in − Com(L1,∞), and hence is strongly measurable. As Ms(H ) is a subspace of L1,∞, it follows that A is strongly measurable if and only if B is strongly measurable. Suppose now that H ′ = H . Let U : H H ′ be a unitary isomorphism, and set B′ = U ∗BU. ′ 6 → Then B is an operator in L1,∞(H ) with same non-zero eigenvalues and same multiplicities as A and B. Thus, by the first part of the proof, A is strongly measurable if and only B′ is strongly measurable. ′ ∗ Let αU : L (H ) L (H ) be defined by αU (T ) = U TU, T L (H ). This is a - ∗ → ∈ ′ ∗ isomorphism of C -algebras. It induces an isometric isomorphism αU : L1,∞(H ) L1,∞(H ). → ′ By duality we get a one-to-one correspondance ϕ ϕ αU between positive traces on L1,∞(H ) → ◦ ′ and positive traces on L1,∞(H ). It then follows that B is strongly measurable if and if only B is strongly measurable. This gives the result when H ′ = H . The proof is complete.  6 Using the same the same type of argument that lead to Corollary 2.23 and Corollary 2.22 we obtain the following consequence of Proposition 4.9. Corollary 4.10. Let ι : H ′ H be a Hilbert space embedding. → ′ (1) If A L1,∞(H ), then ι∗A is strongly measurable if and if only A is strongly measurable. ∈ ′ −1 (2) If ι is an isomorphism, then ιMs(H )ι = Ms(H ). 4.2. Connes’ trace theorem. Suppose that (M n,g) is a closed Riemannian manifold and E is a Hermitian vector bundle. Given m R, we denote by Ψm(M, E) the space of m-th order classical pseudodifferential operators (ΨDOs)∈ P : C∞(M, E) C∞(M, E). If P Ψm(M, E), then we denote by σ(P )(x, ξ) its principal symbol; this is a smooth→ section of End(E∈) over T ∗M 0. Any m \ P Ψ (M, E) with m 0 extends to a bounded operator P : L2(M, E) L2(M, E). If in ∈ ≤ → −1 addition m< 0, then we get an operator in the weak Schatten class Lp,∞ with p = n m , i.e., an infinitesimal operator of order 1/p. | | Z m Z Setting Ψ (M, E)= m∈Z Ψ (M, E), let Res : Ψ (M, E) C be the noncommutative residue trace of Guillemin [28] and Wodzicki [65]. It appears as the residual→ trace on integer-order ΨDOs induced by the analyticS extension of the ordinary trace to all non-integer order ΨDOs (see [28, 65]). A result of Wodzicki [64] further asserts this is the unique trace up to constant multiple on the algebra ΨZ(M, E) if M is connected and has dimension n 2 (see also [32, 43]). The noncommutative residue is a local functional. Namely, if P ΨZ(M,≥ E), then ∈

Res(P )= trE[cP (x)], ZM where cP (x) is an End(E)-valued 1-density which is given in local coordinates by

−n n−1 cP (x)=(2π) a−n(x, ξ)d ξ, Z|ξ|=1 where a (x, ξ) is the symbol of degree n of P . In particular, if P has order n, then −n − − −n Res(P )=(2π) trE σ(P )(x, ξ) dxdξ, ∗ ZS M ∗ ∗ ∗   where S M = T M/R+ is the cosphere bundle and dxdξ is the Liouville measure. Proposition 4.11 (Connes’s Trace Theorem [17, 29]). Every operator P Ψ−n(M, E) is strongly measurable, and we have ∈ 1 (4.2) P = Res(P ). − n Z Remark 4.12. Connes [17] established measurablity and derived the trace formula 4.2. Kalton- Lord-Potapov-Sukochev [29] obtained strong measurability. We observe that Connes’ arguments can also be used to get strong measurability. 16 Suppose that E is the trivial line bundle, and let ∆g be the Laplace-Beltrami operator on functions. As an application of Connes’ trace theorem we obtain the following integration formula, which shows that the noncommutative integral recaptures the Riemannian volume density. Proposition 4.13 (Connes’s Integration Formula [17, 29]). For all f C∞(M), the operator n − 2 ∈ f∆g is strongly measurable, and we have

n n − 4 − 4 n 1 −n n−1 (4.3) ∆g f∆g = cn f(x) det (g(x))d x, cn := (2π) S . − M n | | Z Z p Remark 4.14. The integration formula (4.3) fails in general for functions in L1(M) (see [29]). However, as shown by Rozenblum [49] and Sukochev-Zanin [61] (see also [45]) it actually holds for any function in the LlogL(M), i.e., measurable functions f on M such that (1 + f ) log(1 + f )√gdx < . In particular, it holds for any f Lp(M) with p > 1 (see also [29| , |33, 35]).| In| fact, Rozenblum∞ [49] further extends this result∈ to potentials of the form R f = hµ, where h LlogL(M) and µ is in a suitable class of Borel measures (see also [50]). ∈ Remark 4.15. We refer to [42] for versions of Connes’ trace theorem and Conne’s integration formulas for Heisenberg pseudodifferential operators on contact manifolds and Cauchy-Riemann manifolds. We also refer to [36, 37, 44] for extensions of Connes’ trace theorem and Connes’s integration formula to noncommutative tori. In addition, versions of Connes’ integration formula for noncommutative Euclidean spaces and SU(2) are given in [38].

5. Weyl’s Laws and Noncommutative Integration In this section, we relate Connes’ integration to the Weyl’s laws for compact operators studied by Birman-Solomyak [7] and others in the late 60s and early 70s. In particular, this will exhibit an even stronger notion of measurability of purely spectral nature, and so this will provide another spectral theoretic interpretation of Connes’ integral. 5.1. Weyl operators. If A is a selfadjoint compact operator, then as in Remark 2.4 we denote by ± ± ± ± ( λ (A))j≥0 its sequences of positive and negative eigenvalues, i.e., λj (A)= λj (A )= µj (A ), ± ± 1 where A = 2 ( A A) are the positive and negative parts of A. We refer to [15, §9.2] for the main properties| of|± the positive/negative eigenvalue sequences of selfadjoint compact operators. In particular, we have the following min-max principle (cf.[15, Theorem 9.2.4]),

± Aξ ξ λj (A) = min max h | i ; dim E = j , j 0. 06=ξ∈E⊥ ± ξ ξ ≥  h | i  This implies the following version of Ky Fan’s inequality (cf.[15, Theorem 9.2.8]), (5.1) λ± (A + B) λ±(A)+ λ±(B), j, k 0. j+k ≤ j k ≥ Definition 5.1. We say that A Lp,∞, p> 0, is a Weyl operator if one of the following conditions applies: ∈ 1/p (i) A 0 and lim j λj (A) exists. ∗≥ 1/p + 1/p − (ii) A = A and lim j λj (A) and lim j λj (A) both exist. 1 ∗ 1 ∗ (iii) The real part A = 2 (A + A ) and the imaginary part A = 2i (A A ) of are both Weyl operators in theℜ sense of (ii). ℑ − We denote by W the class of Weyl operators in L . If A W , A 0, we set p,∞ p,∞ ∈ p,∞ ≥ 1 Λ(A) = lim j p λj (A). j→∞ ∗ If A = A Wp,∞, we set ∈ 1 ± Λ±(A) = lim j p λ (A). j→∞ j For an arbitrary operator A W , we define ∈ p,∞ Λ±(A)=Λ± A + iΛ± A . ℜ ℑ 17   Remark 5.2. If A = A∗ L , then ∈ p,∞ − 1 (5.2) 0 λ±(A) µ (A) (j + 1) p A j 0. ≤ j ≤ j ≤ k kp,∞ ∀ ≥ 1/p ± In particular, the sequences (j λ (A))j≥0 are always bounded. If in addition A is a Weyl operator, then we have (5.3) 0 Λ±(A) A . ≤ ≤k kp,∞ ∗ 1/p Remark 5.3. If A = A (Lp,∞)0, then j µj (A) 0, and so by using (5.2) we see that 1/p ± ∈ W ± → j λj (A) 0 as well. Thus, A p,∞, and Λ (A) = 0. More generally, by taking real and → ∈ ± imaginary parts we see that every operator A (Lp,∞)0 is contained in Wp,∞ with Λ (A) = 0. This includes all infinitesimal operators of order∈ >p. Remark 5.4. Given any selfadjoint compact operator A on H , its counting functions are given by ± ± N (A; λ) := # j; λj (A) > λ , λ> 0.

If A Wp,∞, p> 0, then (see, e.g., [57, Proposition 13.1]), we have ∈ p ± ± p ± p (5.4) lim λ N (A; λ) = lim j(λj (A)) =Λ (A) . λ→0+ j→∞ In addition, it will be convenient to introduce the following class of operators.

Definition 5.5. W|p,∞|, p > 0, consists of operators A Lp,∞ such that A Wp,∞, i.e., 1/p ∈ | | ∈ lim j µj (A) exists. In particular, if A W , then ∈ |p,∞| 1 Λ A = lim j p µj (A). | | j→∞ 5.2. Birman-Solomyak’s perturbation  theory. We recall the main facts regarding the per- turbation theory of Birman-Solomyak [7, §4]. ∗ Proposition 5.6 (Birman-Solomyak [7, Theorem 4.1]). Let A = A Lp,∞. Assume that, for every ǫ> 0, we may write ∈ ′ ′′ A = Aǫ + Aǫ , where A′ and A′′ are selfadjoint operators in L such that A′ W , and ǫ ǫ p,∞ ǫ ∈ p,∞ 1 ± lim sup j p λ (A′′) ǫ. j ǫ ≤ Then A Wp,∞, and we have ∈ ± ′ ± lim Λ Aǫ =Λ (A). ǫ→0+ We stress that Proposition 5.6 is obtained in [7] as a sole consequence of the Ky Fan’s inequal- ity (5.1). This result has a number of consequences. Corollary 5.7. W is a closed subset of L on which Λ± : W C are continuous maps. p,∞ p,∞ p,∞ → Proof. We only need to show that if (Aℓ)ℓ≥0 is a sequence in Wp,∞ converging to A in Lp,∞, then A W , and we have ∈ p,∞ ± ± (5.5) lim Λ (Aℓ)=Λ (A). ℓ→∞

By taking real and imaginary parts we may assume that the operators Aℓ and A are selfadjoint. Thanks to (5.3) we have sup Λ±(A ) sup A < . That is, Λ±(A ) are ℓ≥0 ℓ ≤ ℓ≥0 k ℓkp,∞ ∞ { ℓ }ℓ≥0 bounded sequences. Let Λ±(A ) be convergent subsequences. As A A in L , given { ℓp }p≥0 ℓp → p,∞ any ǫ> 0, there is p >ǫ−1 such that A A <ǫ. In view of (5.2) this implies that ǫ k − ℓpǫ k 1 ± lim sup j p λ A A A A < ǫ. j − ℓpǫ ≤ − ℓpǫ W W As Aℓpǫ p,∞ for all ǫ> 0, it follows from Proposition 5.6 that A p,∞ and we have ∈ ∈ ± ± Λ(A) = lim Λ Aℓp = lim Λ Aℓp . ǫ→0+ ǫ p→∞ 18   ± ± This shows that Λ (A) are the unique limit points of the bounded sequences Λ (Aℓ) ℓ≥0. This gives (5.5). The proof is complete. { } 

Corollary 5.8 (K. Fan [25, Theorem 3]; see also [26, Theorem II.2.3]). Let A Wp,∞ and B (L ) . Then A + B W , and we have ∈ ∈ p,∞ 0 ∈ p,∞ Λ±(A + B)=Λ±(A). Proof. It is enough to prove the result when A and B are selfadjoint. In this case we know from 1/p ± Remark 5.3 that lim j λj (B) = 0. Therefore, the we get the result by applying Proposition 5.6 ′ ′′  to A + B with Aǫ = A and Aǫ = B for all ǫ> 0. As mentioned above Proposition 5.6 is a sole consequence of the Ky Fan’s inequality (5.1). As the singular values satisfy the Ky Fan’s inequality (2.3), we similarly have a version of Proposi- tion 5.6 for singular values (cf.[7, Remark 4.2]). Namely, we have the following result.

Proposition 5.9 (Birman-Solomyak [7]). Let A Lp,∞. Assume that, for every ǫ> 0, we may write ∈ ′ ′′ A = Aǫ + Aǫ , with A′ W and A′′ L such that ǫ ∈ |p,∞| ǫ ∈ p,∞ 1 lim sup j p µ (A′′) ǫ. j ǫ ≤ Then A W|p,∞|, and we have ∈ ′ lim Λ Aǫ = Λ( A ). ǫ→0+ | | | | By arguing along the same lines as that of the proofs of Corollary 5.7 and Corollary 5.8 we arrive at the following statements.

Corollary 5.10. W|p,∞| is a closed subset of Lp,∞ on which the functional A Λ( A ) is con- tinuous. → | |

Corollary 5.11 (K. Fan [25, Theorem 3]; see also [26, Theorem II.2.3]). If A W|p,∞| and B (L ) . Then A + B W , and we have ∈ ∈ p,∞ 0 ∈ |p,∞| Λ± A + B =Λ± A . | | | | 5.3. Measurability of Weyl Operators. Suppose that p = 1. We shall now show that every Weyl operator in L1,∞ is strongly measurable and explain how to compute its NC integral in terms of the maps Λ±. More precisely, we shall prove the following result. Proposition 5.12. Let A W . Then A is strongly measurable, and we have ∈ 1,∞ (5.6) A =Λ+(A) Λ−(A). − − Z In particular, if A is selfadjoint, then

A = lim jλ+(A) lim jλ−(A). − j→∞ j − j→∞ j Z Proof. Let us first show that A is measurable and its integral is given by (5.6). By taking real and imaginary parts and using Proposition 2.19 we may assume that A is selfadjoint. In this case we have

1 + − + − + − lim (λj (A) λj (A)) = lim jλj (A) lim jλj (A)=Λ (A) Λ (A). N→∞ log N − j→∞ − j→∞ − j

This implies that jµj (B) sup j γk sup k γk . Thus, ≤ k≥j | |≤ k≥j | | lim sup jµj (B) lim sup k γk = lim sup j γj = lim j γj =0. j→∞ ≤ j→∞ k≥j | | j→∞ | | j→∞ | |

This shows that B is in (L1,∞)0, and hence is strongly measurable by Proposition 4.5. As A = B + Λ(T )T0, it follows that A is strongly measurable as well. The proof is complete.  Corollary 5.13. Let A W . Then A is strongly measurable, and we have ∈ |1,∞| | |

A = lim jµj(A) − | | j→∞ Z 6. Weyl’s Laws for Negative Order ΨDOs In the 70s Birman-Solomyak [10, 11, 12] established a Weyl’s law for negative order ΨDOs. Unfortunately, the main key technical details are exposed in a somewhat compressed manner in the Russian article [11], the translation of which remains unavailable. In this section, after explaining how this implies a stronger version of Connes’s trace theorem, we shall provide a “soft proof” of Birman-Solomyak’s result. This will answer Question D. 6.1. Weyl’s law for negative order ΨDOs. In the following (M n,g) is a closed Riemannian manifold and E is a Hermitian vector bundle over M. We keep on using the notation of §§4.2. Theorem 6.1 (Birman-Solomyak [10, 11, 12]). Let P Ψ−m(M, E), m< 0, and set p = nm−1. ∈ (1) P and P are Weyl operators in Lp,∞. (2) We have| | 1 p 1 1 −n p lim j p µj (P )= (2π) trE σ(P )(x, ξ) dxdξ . j→∞ n ∗ | |  ZS M  (3) If P is selfadjoint, then   1 1 1 p p ± −n p lim j λj (P )= (2π) trE σ(P )(x, ξ)± dxdξ . j→∞ n ∗  ZS M    Remark 6.2. In [10, 11] Birman-Solomyak established the Weyl’s laws above for compactly sup- ported pseudodifferential operators on Rn under very low regularity assumptions on the symbols. Furthermore, the symbols are allowed to be anisotropic. This was extended to classical ΨDOs on closed manifolds in [12]. Remark 6.3. We refer to [1, 3, 16, 22, 27, 31, 49, 50], and the references therein, for various generalizations and applications of Birman-Solomyak’s asymptotics. Combining Theorem 6.1 with Corollary 5.12 and Corollary 5.13 provides us with a stronger form of Connes’ trace theorem (compare Proposition 4.11). Theorem 6.4. Let P Ψ−n(M, E). The following holds. ∈ (1) P and P are Weyl operators in L1,∞, and hence are strongly measurable. (2) We have| |

1 −n (6.1) P = Λ(P )= (2π) trE σ(P )(x, ξ) dxdξ, − n ∗ Z ZS M 1 −n   (6.2) P = lim jµj(P )= (2π) trE σ(P )(x, ξ) dxdξ. − | | j→∞ n ∗ | | Z ZS M 20   (3) If P is selfadjoint, then

± 1 −n (6.3) lim jλj (P )= (2π) trE σ(P )(x, ξ)± dxdξ. j→∞ n ∗ ZS M ∗   In what follows we let ∆E = be the Laplacian of some Hermitian connection on E. ∇ ∇ 2 ∇ In particular, ∆E is (formally) selfadjoint and has principal symbol σ(∆E )= ξ idEx (where we denote by the Riemannian metric on T ∗M). | | Theorem| ·6.1 | leads to the following improvement of Connes’ integration formula (4.3).

Corollary 6.5. Let u C∞(M, End(E)). The following hold. ∈ E −n/4 E −n/4 E −n/4 E −n/4 (1) The operators (∆ ) u(∆ ) and (∆ ) u(∆ ) are Weyl operators in L1,∞, and hence are strongly measurable. | | (2) We have

− n − n 1 (6.4) ∆ 4 u∆ 4 = (2π)−n tr u(x) g(x)dx, − E E n E Z ZM n n n n  p − 4 − 4 − 4 − 4 1 −n (6.5) ∆E u∆E = lim µj ∆E u∆E = (2π) trE u(x) g(x)dx. − j→∞ n M | | Z   Z ∗  p (3) If u( x) = u(x), then

n n ± − 4 − 4 1 −n (6.6) lim jλ ∆ u∆ = (2π) trE u(x)± g(x)dx. j→∞ j E E n ZM    p Remark 6.6. The above results continue to hold for operators of the form QuP , where P and Q are operators in Ψ−n/2(M, E) and u(x) is a potential in the Orlicz class LlogL(M, End(E)) (for the 3rd part we take Q = P ∗) (see [49, 45]; see also [61]). In particular, the operators P and Q need not be a negative power of an elliptic operator. Rozenblum [49] actually establishes the results in the scalar case for potentials of the form u = hµ, where h is in the Orlicz space LlogL(M) and µ is in a suitable class of Borel measures (see also [50]). If in addition, if E is a Clifford module, then 2 we may replace ∆E by D/ E, where D/ E is the Dirac operator associated to some unitary Clifford connection on E.

6.2. Proof of Theorem 6.1. We will deduce Theorem 6.1 from the properties of zeta functions of elliptic operators and their relationship with the noncommutative residue trace. This will clarify the relationship between Birman-Solomyak’s result and the noncommutative residue. More precisely, we shall use the following result.

Proposition 6.7 ([28, 65]). Let Q Ψm(M, E), m > 0, be elliptic and let A Ψ0(M, E). The function z Tr[A Q −z] has a meromorphic∈ extension to C with at worst simple∈ pole singularities on Σ := km→−1; k| |Z, k n in such way that { ∈ ≤ } 1 Res Tr A Q −z = Res A Q −σ , σ Σ. z=σ | | m | | ∈ As is well known, the above result implies the following Weyl’s laws.

Corollary 6.8. Let Q Ψm(M, E), m> 0, be elliptic. Set p = nm−1. ∈ (1) We have

1 p 1 −1 1 −n −p (6.7) lim j p µj Q = (2π) trE σ(Q)(x, ξ) dxdξ . j→∞ n ∗ | |  ZS M     (2) If Q is selfadjoint, then

1 1 1 p p ± −1 −n −p (6.8) lim j λj Q = (2π) trE σ(Q)(x, ξ)± dxdξ . j→∞ n ∗  ZS M   21   Proof. The first part is a mere restatement of the Weyl’s law for Q . Namely, Proposition 6.7 for A = 1 implies that the function Tr[ Q −s]= µ ( Q −1)s has| a meromorphic| extension to the | | j≥0 j | | half-space s>p 1/m with a single pole at s = p such that ℜ − P 1 −p 1 −n −p Res Q = (2π) trE σ(Q)(x, ξ) dxdξ. m | | m ∗ | | ZS M By using Ikehara’s Tauberian theorem (see, e.g., [57]) we then obtain the Weyl’s law (6.7). Suppose now that Q is selfadjoint. Let Π0(Q) be the orthogonal projections onto ker Q and Π±(Q) the orthogonal projections onto the positive and negative eigenspaces of Q. Here Π0(Q) is a smoothing operator, and Π (Q) are ΨDOs of order 0, since ± ≤ 1 Π (Q)= 1 Π (Q) Q Q −1 . ± 2 − 0 ± | | In particular, σ(Π (Q)) = 1 (1 σ(Q)σ(Q −1)) = Π (σ(Q)). Therefore, Proposition 6.7 for ± 2 ± | | ± A = Π (Q) shows that the function Tr[Π (Q) Q −s] = λ±(Q−1)s has a meromorphic ± ± | | j≥0 j extension to the half-space s>p 1/m with a single pole at s = p such that ℜ − P 1 −p 1 −n −p Res Π±(Q) Q = (2π) trE Π±(σ(Q)) σ(Q)(x, ξ) dxdξ m | | m ∗ | | ZS M  1 −n  −p  = (2π) trE σ(Q)(x, ξ)± dxdξ. m ∗ ZS M As above, using Ikehara’s Tauberian theorem gives (6.8). The proof is complete.  We will also need the following special case of the Birman-Koplienko-Solomyak inequality. Lemma 6.9 ([6, Theorem 3]; see also [14, Proposition 4.9]). Let A and B be non-negative selfad- joint operators on H such that A B L , q> 0. Then √A √B L , and − ∈ q,∞ − ∈ 2q,∞ √A √B C A B , − 2q,∞ ≤ q k − kq,∞ where the constant C depends only on q. q q We are now in a position to prove Theorem 6.1. Proof of Theorem 6.1. Let P Ψm(M, E), m > 0, and set p = nm−1. Throughout this proof ∗ ∈ we let ∆E = be the Laplacian of some Hermitian connection on E. In particular, σ(∆E )= ∇−∇m ξ 2 id and ∆ Ψ−2m(M, E). Given ǫ> 0 set | | Ex E ∈ ∗ 2 −m Aǫ = P P + ǫ ∆E . 2 ∗ 2 −m L q L Here Aǫ P P = ǫ ∆E p/2,∞. Therefore, by Lemma 6.9 the difference Aǫ P is in p,∞, and we have− − ∈ −| | A P C ǫ ∆−m , ǫ − | | p,∞ ≤ p E p/2,∞ where the constant Cp does not depend on ǫ. Thus,q

(6.9) A P in L as ǫ 0. ǫ −→ | | p,∞ → −1/2 Let Q Ψm(M, E) have principal symbol σ(P )(x, ξ) 2 + ǫ2 ξ −2m = σ(A2)−1/2. In ǫ ∈ | | | | ǫ particular, Q is an elliptic operator. Thus, by Corollary 6.8 the inverse absolute value Q −1 is ǫ  | ǫ| a Weyl operator in Lp,∞, and we have 1 p p −1 1 −n 2 2 −2m 2 Λ Qǫ = (2π) trE σ(P )(x, ξ) + ǫ ξ dxdξ | | n S∗M | | | |  Z 1    p   1 −n p (2π) trE σ(P )(x, ξ) dxdξ . −−−→ǫ→0 n ∗ | |  ZS M  −2 2 2 −2  −2m−1 By construction σ( Qǫ )= σ(Aǫ ), so Aǫ Qǫ is an operator in Ψ (M, E), and hence | | −| | −1 is in the weak Schatten class Lq/2,∞ with q = 2n(2m + 1) < p. Lemma 6.9 then ensures us 22 that A Q −1 is in the weak Schatten class L , and hence is contained in (L ) . It then ǫ − | ǫ| q,∞ p,∞ 0 follows from Corollary 5.11 that Aǫ is a Weyl operator in Lp,∞, and we have

1 p −1 1 −n p Λ(Aǫ)=Λ Qǫ (2π) trE σ(P )(x, ξ) dxdξ . | | −−−→ǫ→0 n ∗ | |  ZS M     Combining this with Corollary 5.10 and (6.9) then shows that P is a Weyl operator in Lp,∞, and we have | | 1 p 1 −n p Λ( P ) = lim Λ(Aǫ)= (2π) trE σ(P )(x, ξ) dxdξ . | | ǫ→0 n ∗ | |  ZS M  ∗ 1   Suppose now that P = P . Set Bǫ = 2 (Aǫ + P ). It follows from (6.9) that 1 (6.10) Bǫ ( P + P )= P+ in Lp,∞. −−−→ǫ→0 2 | | In addition, let Q˜ Ψm(M, E) be selfadjoint and have principal symbol ǫ ∈ 1 1 −1 1 1 −1 σ(P )(x, ξ) 2 + ǫ2 ξ −2m + σ(P )(x, ξ) = σ Q −1 + σ(P )(x, ξ) . 2 | | | | 2 2 | ǫ| 2     p ˜ ˜−1 L As Qǫ is elliptic, Corollary 6.8 ensures that Qǫ is a Weyl operator in p,∞, and we have p 1 1 1 1 p + ˜−1 −n 2 2 −2m Λ Qǫ = (2π) trE σ(P )(x, ξ) + ǫ ξ + σ(P )(x, ξ) dxdξ . n ∗ 2 | | | | 2  ZS M     In particular,  p 1 p p + ˜−1 1 −n 1 1 lim Λ Qǫ = (2π) trE σ(P )(x, ξ) + σ(P )(x, ξ) dxdξ ǫ→0 n S∗M 2| | 2  Z  1     p 1 −n p = (2π) trE σ(P )(x, ξ)+ dxdξ . n ∗  ZS M  ˜−1 1 −1   ˜−1 1 −1 By construction, σ(Qǫ ) = 2 (σ( Qǫ )+ σ(P )). Thus, Qǫ 2 ( Qǫ + P ) is a ΨDO of | | − −| 1 | order (m + 1), and hence is contained in Lq,˜ ∞ withq ˜ = n(m + 1) . Asq

1 p + + ˜−1 1 −n p Λ (Bǫ)=Λ Qǫ (2π) trE σ(P )(x, ξ)+ dxdξ . −−−→ǫ→0 n ∗  ZS M     Combining this with (6.10) and using Proposition 6.1 shows that P+ is a a Weyl operator in Lp,∞, and we have 1 p + 1 −n p Λ(P+) = lim Λ (Bǫ)= (2π) trE σ(P )(x, ξ)+ dxdξ . ǫ→0 n ∗  ZS M  Upon replacing P by P further shows that P is a Weyl operator in L and Λ(P ) is given − − p,∞ − by (3). This shows that P is a Weyl operator in Lp,∞. The proof of Theorem 6.1 is complete.  Remark 6.10. A similar approach allows us to obtain a Weyl’s law for negative order ΨDOs on NC tori (see [46]). Remark 6.11. As mentioned in Remark 6.2, the original version of Birman-Solomyak’s result on Rn in [10] was established for ΨDOs associated with anisotropic symbols. Therefore, we may expect to have a version of Birman-Solomyak’s result for the Heisenberg calculus [4, 62] and more generally for the pseudodifferential calculus on filtered manifolds [40]. In those settings the pseudodifferential operators are defined in terms of anisotropic symbols. Note that we already have a noncommutative residue trace for the Heisenberg calculus (see [42]). 23 7. Connes’ Integration and Semiclassical Analysis In this section, we look at the relationship between Connes’ integration and semiclassical Weyl’s laws for abstract Schr¨odinger operators, i.e., we shall answer Question E. This will follow from the Birman-Schwinger principle. Once again our aim is to stress out how general and simple this relationship is (compare [39]). In what follows we let H be a (densely defined) selfadjoint operator on H with non-negative 1 spectrum containing 0. Its quadratic form QH has domain dom(QH ) = dom(H + 1) 2 . We denote by H+ the Hilbert space obtained by endowing dom(QH ) with the Hilbert space norm, 1 ξ = Q (ξ, ξ)+ ξ 2 2 = (1 + H)1/2ξ , ξ dom(Q ). k k+ H k k ∈ H H H We also let − be the Hilbert space of continuous anti-linear functionals on +. Note that we by have a continuous inclusion ι : H ֒ H given → − ι(ξ), η = ξ η , ξ H , η H . h i h | i ∈ ∈ + The operator (H +1) is a unitary isomorphism from H onto H with inverse (H +1)−1/2 : H + → H . By duality we get a unitary isomorphism (H + 1)−1/2 : H H such that + − → (H + 1)−1/2ξ η = ξ, (H + 1)−1/2η , ξ H , η H . | ∈ − ∈ Let V : H H be a bounded operator. We denote by Q the corresponding quadratic + → − V form with domain H+ and given by Q (ξ, η) := Vξ,η , ξ,η H . V h i ∈ + We assume that QV is symmetric and H-form compact. The latter condition means that the −1/2 −1/2 operator V : H+ H− is compact, or equivalently, (H + 1) V (H + 1) is a compact operator on H . → Our main focus is the operator HV := H + V . It makes sense as a bounded operator HV : H+ H−. As the symmetric quadratic form QV is H-form compact, it is H-form bounded with→ zero H-bound (see [60, §7.8]). Therefore, by the KLMN theorem (see, e.g., [48, 51]) the −1 H H restriction of HV to dom(HV ) := HV ( ) is a bounded from below selfadjoint operator on whose quadratic form is precisely QH + QV . It can be further shown that, for all λ Sp(H) Sp(HV ) that H and HV have the same (see, e.g., [60, Theorem6∈ 7.8.4]). Thus,∪ as H has non-negative spectrum, the bottom of the essential spectrum of H is 0. V ≥ As HV is bounded from below, we may list its eigenvalues below the essential spectrum as a non-decreasing sequence, λ (H ) λ (H ) λ (H ) , 0 V ≤ 1 V ≤ 2 V ≤ · · · where each eigenvalue is repeated according to multiplicity. This sequence may be finite or infinite, or even empty. We then introduce the counting function,

(7.1) N(HV ; λ) :=# j; λj (HV ) < λ , λ< inf Spess(HV ). − We also set N (HV )= N(HV ; 0).  Assume further that 0 lies in the discrete spectrum of H, i.e., 0 is an isolated eigenvalue of H. Thus, the essential spectrum of H is contained in some interval [a, ) with a> 0. We denote by −1/2 1/2 ∞ H the partial inverse of H . Moreover, as H and HV have the same essential spectrum, it follows that HV has at most finitely many non-positive eigenvalues. The Birman-Schwinger principle was established by Birman [5] and Schwinger [56] for Schr¨odinger operators ∆+V on Rn, n 3. Its abstract version [13, Lemma 1.4] (see also [36, Proposition 7.9]) ≥ allows us to relate the number of negative eigenvalues of HV to the counting functions of the 1 1 2 Birman-Schwinger operator H− 2 VH− / . Note also that the assumptions on V ensure us that 1 1 H− 2 VH− 2 is a selfadjoint compact operator. Proposition 7.1 (Abstract Birman-Schwinger Principle [13]; see also [36]). Under the above assumptions, we have

1 1 1 1 (7.2) N − H− 2 VH− 2 ;1 N −(H ) N − H− 2 VH− 2 ;1 + dim ker H. ≤ V ≤   24   Corollary 7.2 (Birman-Solomyak; see [8, Theorem 10.1] and [9, Appendix 6]). Assume further −1/2 −1/2 that H VH is a Weyl operator in Lp,∞ for some p > 0. Then, under the semiclassical limit h 0+, we have → 1 1 p 2p − 2 − − 2 − 2 (7.3) lim h N h H + V = lim jλj H VH . h→0+ j→∞    Proof. The Birman-Schwinger principle (7.2) ensures that, as h 0+, we have → N − h2H + V = N −(H + h−2V ),

1 2 1 (7.4)  = N − H− 2 h− VH− 2 ;1 + O(1),

 1 1 2  = N − H− 2 VH− 2 ; h + O(1).   − 1 − 1 If H 2 VH 2 is a Weyl operator in Lp,∞ for some p> 0, then by Remark 5.4 we have

1 1 1 1 p 2p − − 2 − 2 2 − − 2 − 2 lim h N H VH ; h = lim jλj H VH . h→0+ j→∞     Combining this with (7.4) immediately gives the result. 

Combining the above corollary with Proposition 5.12 leads us to the following semiclassical interpretation of Connes’ integral.

− 1 − 1 Proposition 7.3. Assume that V 0 and H 2 VH 2 is a Weyl operator in L1,∞. Then, the 1 1 ≥ operator H− 2 VH− 2 is strongly measurable, and we have

2 2 1 1 (7.5) lim h N − h H V = H− 2 VH− 2 . h→0+ − − Z  In a recent preprint McDonald-Sukochev-Zanin [39] obtained a related formula in the setting of spectral triples under some additional technical conditions. We stress out that, on the one hand, the equality (7.5) holds in fairly great generality. For instance, we don’t have to require H to have compact resolvent. On the other hand, this equality is an immediate consequence of the abstract Birman-Schwinger principle (compare [39]). In any case, the formula (7.5) highlights a neat link between the semiclassical analysis of Schr¨odinger operators and Connes’ noncommutative geometry. These are two different approaches to quantum theory, and so it is quite interesting to witness some intersection between them. n To illustrate the above results, suppose that (M ,g) be a closed Riemannian manifold. Let ∆g n/2 be the corresponding Laplace-Beltami operator acting on functions, and take H = ∆g . In this H n/2 H −n/2 setup + is the W2 (M) and − is the antilinear dual W2 (M). Let V (x) be a real-valued potential in the Orlicz class LlogL(M), i.e., a measurable function such that

(1 + V (x) ) log(1 + V (x) ) g(x)dx < . | | | | ∞ Z p For instance, we may take V (x) to be in Lp(M) for any p> 1. It can be shown that V (x) gives rise to a bounded operator V : W n/2(M) W −n/2(M) −n/4 −n/4 → (see [49]), and so ∆g V ∆g is bounded on L2(M). As mentioned in Remark 4.14, results of −n/4 −n/4 Rozenblum [49] and Sukochev-Zanin [61] (see also [45, 50]) ensure us that ∆g V ∆g is a Weyl n/2 operator in L1,∞. Thus, ∆g + V makes sense as a form sum as above, and by Corollary 7.2 we have

n n n − n n/2 − − 4 − 4 lim h N h ∆g + V = lim jλj ∆g V ∆g h→0+ j→∞  1   = (2π)−n V (x) g(x)dx. n − ZM 25 p If V (x) 0, then by using Proposition 7.3 we further obtain ≥ n n n − 4 − 4 n − n 2 ∆g V ∆g = lim h N h ∆g V , − h→0+ − Z 1  = (2π)−n V (x) g(x)dx. n ZM This provides us with a semiclassical interpretation of Connes’p integration formula (4.3). (See also [45] for an extension of this result to matrix-valued Orlicz-LlogL potentials.)

Appendix A. Embeddings of Hilbert spaces In this section, we gather a few basic facts about embedding of Hilbert spaces and their actions on Schatten and weak Schatten classes. Given quasi-Banach spaces E and E ′, a continuous linear embedding ι : E ′ E is a continuous linear map which is one-to-one and has closed range. For instance, any isometric→ linear map ι : E ′ E is an embedding. Suppose→ now that H and H ′ are Hilbert spaces, and let ι : H ′ H be a continuous linear → embedding. Denote by H1 the range of ι. By assumption this is a closed subspace of H . The ′ ′ embedding ι : H H induces a continuous linear isomorphism ι : H H1 whose inverse is −1 → ′ → −1 denoted ι : H1 H . Let π : H H be the orthogonal projection onto H1. Then ι π is a left-inverse of →ι. More precisely, → ◦ −1 −1 (ι π) ι = idH ′ , ι (ι π)= π. ◦ ◦ ◦ ◦ The pushforward ι : L (H ′) L (H ) is then defined by ∗ → (A.1) ι A = ι A (ι−1 π), A L (H ′). ∗ ◦ ◦ ◦ ∈ In fact, with respect to the orthogonal splitting H = H H ⊥ we have 1 ⊕ 1 ιAι−1 0 (A.2) ι∗A = . 0 0! ′ In particular, we see that the pushforward map ι∗ : L (H ) L (H ) is a continuous embedding. It is also multiplicative, and so we get an embedding of (unital)→ Banach algebras. ′ If in addition ι is an isometric embedding, then ι∗ : L (H ) L (H ) is an isometric embedding ′ → as well. In this case, we ι : H H1 is a , and so in view of (A.2) we have ∗ ∗ → ′ ι∗(A ) = (ι∗A) . Thus, in this case ι∗ : L (H ) L (H ) even is an (isometric) embedding of C∗-algebras. → ′ ′ In any case the pushforward ι∗ : L (H ) L (H ) maps compact operators on H to compact operators on H . Moreover, in view of (A.2→) we have the following result.

′ Proposition A.1. If A is a compact operator on H , then A and ι∗A have the same non-zero eigenvalues with the same algebraic multiplicities. In particular, any eigenvalue sequence for A is an eigenvalue sequence for ι∗A, and vice versa. ′ ′ Suppose that ι : H H is an isometric embedding. As mentioned above ι∗ : L (H ) L (H ) is an isometric embdedding→ of C∗-algebras. Thus, for any A L (H ′) and f C(Sp(A→)) we have ∈ ∈ ιf(A)ι−1 0 f(ι∗A)= = ι∗f(A). 0 0! In particular, we have ι A = ι A , and so ι A and ι A have the same non-zero eigenvalues | ∗ | ∗| | | ∗ | ∗| | with same multiplicity. This means that A and ι∗A have the same singular value sequence. ′ In general, as ι : H H1 is a continuous linear isomorphism, we can pullback the inner → ′ product on H1 to a new inner product on H , which is equivalent to its original inner product. With respect to this new inner product the embedding ι : H ′ H becomes isometric. Thus, ′ → given any compact operator A on H , the singular value sequence of ι∗A agrees with the singular value sequence of A with respect to the new inner product on H ′. As this inner product is 26 equivalent to the original inner product of H ′, by using the min-max principle (2.1) we eventually arrive at the following result. Proposition A.2. Let ι : H ′ H be a continuous linear embedding. → (1) There is c> 0 such that, for every compact operator A on H ′, we have c−1µ (A) µ ι A) cµ (A) j 0. j ≤ j ∗ ≤ j ∀ ≥ We may take c =1 when ι is an isometric embedding. ′ ′ (2) Given any p > 0, the operator A is in the class Lp(H ) (resp., Lp,∞(H )) if and only if ι∗A is in Lp(H ) (resp., Lp,∞(H )). Moreover, the pushforward map (A.1) induces continuous linear embeddings, ι : L (H ′) L (H ), ι : L (H ′) L (H ). ∗ p −→ p ∗ p,∞ −→ p,∞ These embeddings are isometric whenever ι is an isometric embedding. Remark A.3. Part (ii) holds more generally for any symmetrically normed ideal. Aknowledgements. I wish to thank Alain Connes, Grigori Rozenblum, Edward McDonald, and Fedor Sukochev for various stimulating discussions related to the subject matter of this article. I also thank the University of Ottawa for its hospitality during the whole preparation of this paper.

References

[1] Agranovich, M.S.; Amosov, B.A.: Estimates for s-numbers, and spectral asymptotics for integral operators of potential type on nonsmooth surfaces. Funct. Anal. Appl. 30 (1996), no. 2, 75–89 [2] Andreev, A.S.: An estimate of the remainder in the spectral asymptotics of pseudodijferential operators of negative order. Funct. Anal. Appl. 20 (1986), 124–125. [3] Andreev, A.S.: Asymptotics of the spectrum of a compact pseudodifferential operators in a Euclidean domain. Math. USSR Sb. 65 (1990), 205–228. [4] Beals, R.; Greiner, P.: Calculus on Heisenberg manifolds. Annals of Mathematics Studies, 119, Princeton University Press, Princeton, NJ, 1988. [5] Birman, M.Sh.: The spectrum of singular boundary problems, Amer. Math. Soc. Transl. (2) 53 (1966), 23–80. [6] Birman, M.S.; Koplienko, L.S.; Solomyak, M.Z.: Estimates of the spectrum of a difference of fractional powers of selfadjoint operators. Izv. Vyssh. Uchebn. Zaved. Matematika, 3(154) (1975), 3–10 (Russian). [7] Birman, M.Sh.; Solomyak, M.Z.: The leading term of the spectral asymptotics for “non-smooth” elliptic problems. Funct. Anal. Appl. 4 (1970), 265–275. [8] Birman, M.Sh.; Solomyak, M.Z.: Spectral asymptotics of nonsmooth elliptic operators. I. Trudy Moskov. Mat. Obshch. 27 (1972), 3–52 (Russian). Trans. Moscow Math. Soc. 27 (1972), 1–52 (1975) (English). [9] Birman, M.Sh., Solomyak, M.Z.: Quantitative analysis in Sobolev imbedding theorems and applications to . Izdanie Inst. Mat. Akad. Nauk Ukrain. SSR, Kiev, 1974 (Russian). American Mathematical Society Translations, Series 2, 114. American Mathematical Society, Providence, R.I., 1980. [10] Birman, M.S.; Solomyak, M.Z.: Asymptotic behavior of the spectrum of pseudodifferential operators with anisotropically homogeneous symbols. Vestnik Leningrad. Univ. 13, no. 3 (1977), 13–21 (Russian). Vestn. Leningr. Univ., Math. 10 (1982), 237–247 (English). [11] Birman, M.S.; Solomyak, M.Z.: Asymptotic behavior of the spectrum of pseudodifferential operators with anisotropically homogeneous symbols. II. Vestnik Leningrad. Univ. Mat. Mekh. Astronom. 13, no. 3 (1979), 5–10 (Russian). [12] Birman, M.S., Solomyak, M.Z.: Asymptotics of the spectrum of variational problems on solutions of elliptic equations. Sib. Math J. 20 (1979), 1–15. [13] Birman M.Sh., Solomyak M.Z.: Schr¨odinger operator. Estimates for number of bound states as function- theoretical problem. Spectral theory of operators (Novgorod, 1989), 1–54. Amer. Math. Soc. Transl. Ser. 2, 150, Amer. Math. Soc., Providence, RI, 1992. [14] Birman, M.S.; Solomyak, M.Z.: Estimates for the difference of fractional powers of selfadjoint operators under unbounded perturbations. J. Sov. Math. 61, no. 2 (1992), 2018–2035. [15] Birman, M.Sh., Solomyak, M.Z.: Spectral theory of self-adjoint operators in Hilbert space, D. Reidel, Boston, MA, 1987. [16] Birman, M.S.; Yafaev, D.R.: Asymptotic behavior of the spectrum of the scattering matrix. J. Sov. Math. 25 (1984), 793–814. [17] Connes, A.: The action functional in non-commutative geometry. Comm. Math. Phys. 117 (1988), 673–683. [18] Connes, A.: Noncommutative geometry. Academic Press, Inc., San Diego, CA, 1994. [19] Connes, A.: Noncommutative geometry, the spectral standpoint. New Spaces in Physics, Volume 2, pp. 23–84, Cambridge University Press. 27 [20] Connes, A.; McDonald, E.; Sukochev, F.; Zanin, D.: Conformal trace theorem for Julia sets of quadratic polynomials. Ergodic Theory and Dynamical Systems 39 (2019), 2481–2506. [21] Connes, A.; Moscovici, H.: The local index formula in noncommutative geometry. Geom. Funct. Anal. 5 (1995), 174–243. [22] Dauge, M.; Robert, D.: Weyl’s formula for a class of pseudodifferential operators with negative order on L2(Rn). Pseudodifferential Operators (Oberwolfach, 1986), Lecture Notes in Math., vol. 1256, Springer-Verlag, 1987, pp. 91–122. [23] Dixmier, J.: Existence de traces non normales. C. R. Acad. Sci. Paris S´er. A–B 262 (1966), A1107–A1108. [24] Dykema, K; Figiel, T.; Weiss, G.; Wodzicki, M.: Commutator structure of operator ideals. Adv. Math. 185 (2004), 1–79. [25] Fan, K.: Maximum properties and inequalities for the eigenvalues of completely continuous operators. Proc. Nat. Acad. Sci. U.S.A. 37 (1951), 760–766. [26] Gohberg, I.C.; Krein, M.G.: Introduction to the theory of linear nonselfadjoint operators. Translations of Mathematical Monographs, Vol. 18. American Mathematical Society, Providence, R.I., 1969. [27] Grubb, G.: Spectral asymptotics for nonsmooth singular Green operators. Comm. Partial Differential Equa- tions 39 (2014), no. 3, 530–573. [28] Guillemin, V.: A new proof of Weyl’s formula on the asymptotic distribution of eigenvalues. Adv. Math. 55 (1985), no. 2, 131–160. [29] Kalton, N.; Lord, S.; Potapov, D.; Sukochev, F.: Traces of compact operators and the noncommutative residue. Adv. Math. 235 (2013), 1–55. [30] Krein, S.; Petunin, Ju.; Semenov, E.: Interpolation of Linear Operators. Translations of Mathematical Mono- graphs, Vol. 54. American Mathematical Society, Providence, R.I., 1982. [31] Ivrii, V.: Microlocal Analysis, Sharp Spectral Asymptotics and Applications II. Functional Methods and Eigenvalue Asymptotics. Springer, Cham, 2019. [32] Lesch, M.: On the noncommutative residue for pseudodifferential operators with log-polyhomogeneous symbols. Ann. Global Anal. Geom. 17 (1999), no. 2, 151–187. [33] Lord, S.; Potapov, D.; Sukochev, F.: Measures from Dixmier traces and zeta functions. J. Funct. Anal. 259 (2010), 1915–1949. [34] Lord, S.; Sukochev, F.; Zanin, D.: Singular traces: theory and applications. De Gruyter Studies in Mathe- matics, Vol. 46, Walter de Gruyter, 2012, Theory and applications. [35] Lord, S.; Sukochev, F.; Zanin, D.: A last theorem of Kalton and finiteness of Connes’ integral. J. Funct. Anal. 279 (2020), 108664, 54 pp.. [36] McDonald, E; Ponge, R.: Cwikel estimates and negative eigenvalues of Schr¨odinger operators on noncom- mutative tori. Preprint arXiv:2102.12021, 28 pp.. [37] McDonald, E.; Ponge: Dixmier trace formulas and negative eigenvalues of Schr¨odinger operators on curved noncommutative tori. Preprint arXiv:2103.16869, 32 pp.. [38] McDonald, E.; Sukochev, F.; Zanin, D.: A C∗−algebraic approach to the principal symbol II. Math. Ann. 374 (2019), 273–322. [39] McDonald, E.; Sukochev, F.; Zanin, D.: Semiclassical Weyl law and exact spectral asymptotics in noncom- mutative geometry. Preprint arXiv:2106.02235, 31 pp.. [40] Melin, A.: Lie filtrations and pseudo-differential operators. Preprint, 1982. [41] Meyer, P.A.: Limites m´ediales, d’apr`es Mokobodzki. S´eminaire de probabilit´es VII (Strasbourg, 1971-1972), pp. 198–204, Lecture Notes in Math., 321, Springer, Berlin, 1973. [42] Ponge, R.: Noncommutative residue for Heisenberg manifolds and applications in CR and contact geometry. J. Funct. Anal. 252 (2007), 399–463. [43] Ponge, R.: Traces on pseudodifferential operators and sums of commutators. J. Anal. Math. 110 (2010), 1–30. [44] Ponge, R.: Connes’s trace Theorem for curved noncommutative tori. Application to scalar curvature. J. Math. Phys. 61 (2020), 042301, 27 pp.. [45] Ponge, R.: Weyl’s laws and Connes’ integration formula for matrix-valued LlogL-Orlicz potentials. Preprint, arXiv, July 2021. [46] Ponge, R.: Zeta functions and Weyl’s laws on noncommutative tori. In preparation. [47] Reed, M.; Simon, B.: Methods of modern mathematical physics. I. . Second edition. Academic Press, Inc., New York, 1980. xv+400 pp.. [48] Reed, M.; Simon, B.: Methods of modern mathematical physics. II. Fourier Analysis, Selfadjointness. Aca- demic Press, Inc., New York, 1975. xv+361 pp.. [49] Rozenblum, G.: Eigenvalues of singular measures and Connes noncommutative integration. Preprint arXiv:2103.02067, 33 pp.. [50] Rozenblum, G; Shargorodsky, E.: Eigenvalue estimates and asymptotics for weighted pseudodifferential oper- ators with singular measures in the critical case. Partial differential equations, spectral theory, and mathemat- ical Physics. The Ari Laptev anniversary volume, EMS Series of Congress Reports, Vol. 18, EMS Publishing, 2021, pp. 331–353. [51] Schm¨udgen, K.: Unbounded self-adjoint operators on Hilbert space. Graduate Texts in Mathematics, 265. Springer, Dordrecht, 2012. xx+432 pp.

28 [52] Sedaev, A.; Sukochev, F.: Dixmier measurability in Marcinkiewicz spaces and applications. J. Funct. Anal. 265 (2013), no. 12, 3053–3066. [53] Seeley, R.T.: Complex powers of an elliptic operator. Proc. Sympos. Pure Math., Vol. X, pp. 288–307. Amer- ican Mathematical Society, Providence, RI, 1967. [54] Semenov, E. M.; Sukochev, F. A.; Usachev, A. S.: Geometry of Banach limits and their applications. Russian Math. Surveys 75 (2020), no. 4, 725–763. [55] Semenov, E.; Sukochev, F.; Usachev, A.; Zanin, D.: Banach limits and traces on L1,∞. Adv. Math. 285 (2015), 568–628. [56] Schwinger, J.: On the bound states of a given potential. Proc. Nat. Acad. Sci. U.S.A. 47 (1961), 122–129. [57] Shubin, M.A.: Pseudodifferential operators and spectral theory, 2nd edition. Springer, Berlin, 2001. [58] Simon, B.: Analysis with weak trace ideals and the number of bound states of Schr¨odinger operators. Trans. Amer. Math. Soc. 224 (1976), 367–380. [59] Simon, B.: Trace ideals and their applications. Second edition. Mathematical Surveys and Monographs, 120. American Mathematical Society, Providence, RI, 2005. [60] Simon, B.: Operator Theory: A Comprehensive Course in Analysis, Part 4. American Mathematical Society, Providence, RI, 2015. [61] Sukochev, F.; Zanin, D.: Connes Integration Formula without singular traces. Preprint arXiv:2103.08817, 19 pp.. [62] Taylor, M.E.: Noncommutative microlocal analysis. I. Mem. Amer. Math. Soc. 52 (1984), no. 313, iv+182 pp.. [63] van Erp, E.; Yuncken, R.: A groupoid approach to pseudodifferential calculi. J. Reine Angew. Math. (2019), 151–182. [64] Wodzicki, M.: and noncommutative residue (in Russian). Habilitation Thesis, Steklov Institute, Soviet Academy of Sciences, Moscow, 1984. [65] Wodzicki, M.: Noncommutative residue. I. Fundamentals. K-theory, arithmetic and geometry (Moscow, 1984– 1986), Lecture Notes in Math., Vol. 1289, Springer, Berlin, 1987, pp. 320–399.

School of Mathematics, Sichuan University, Chengdu, China Email address: [email protected]

29