arXiv:2011.11702v4 [math.RA] 16 Jun 2021 position: nessae tews.W fe eoemultiplication denote often We otherwise. stated unless 2 73,17C27. 17A36, element unit multiplicative a with necessarily u hr utpiaini endby defined is where but denote ocmuaieJordan. noncommutative 12 ( yields (1.2) axiom flexible the Linearizing ambiguity. 11 ( (1.1) entos1.1. Definitions suethat assume omttv theory. commutative hogotti paper this Throughout 2010 Date e od n phrases. and words Key naflxbeagbaw write we algebra flexible a In ∗ omttv lersaeflxbe since flexible, are algebras Commutative esuyegnauso idempotents of eigenvalues study We h rtato a upre yteIre cec Foundati Science Israel the by supported was author first The 1 h algebra The (1) (2) LXBEIEPTNSI NONASSOCIATIVE IN IDEMPOTENTS FLEXIBLE ue1,2021. 17, June : ahmtc ujc Classification. Subject fa xsi h ocmuaiestigadacmlt info accumulate th and setting define axes. noncommutative We of idem the pairs flexible in about case. for axis power-associative rules an the fusion of in establish We Albert following idempotent. an of Abstract. uaie o each for mutative) A xy is . power-associative A A ( sflxbe u betv en ognrlz h well-know the generalize to being objective our flexible, is ◦ ) C Fso ue”aelw fmlilcto mn eigenspace among multiplication of laws are rules” “Fusion steagbawt h aevco pc tutr as structure space vector same the with algebra the is ( A A = ) OI OE OVSEGEV YOAV ROWEN LOUIS is xy xs eil ler,pwrascaie uin idempo fusion, power-associative, algebra, flexible axis, flexible { ) 1. z x A xy x ( + ∈ ∈ sa ler ntncsaiyascaie not associative, necessarily (not algebra an is ) Introduction x ALGEBRAS A A. zy ( = xyx fi aifisteiett ( identity the satisfies it if | ) if xy x yx F = 1 ihu aetee,sneteei no is there since parentheses, without = [ rmr:1A5 71,1A0;Secondary: ; 17A20 17A15, 17A05, Primary: ) x x x x sascaie(n hrfr com- therefore (and associative is ] yx, a ( = yz ◦ ∈ y x + ) A, o all for ( 1 = yx vrafield a over ) ntesii f[] eoften We [A]. of spirit the in z 1 2 ) ( ( . yx xy y ) . ∈ + A yx ngat1623/16. grant on } x ) . F . · xy y hogotwe Throughout notion e ihchar( with potents, rmation ) in x A = yjuxta- by x s ( yx F ) tent, . ) A n 6= , 2 LOUIS ROWEN, YOAV SEGEV

In §2 we also often assume that A is power-associative, relying heavily on Albert [A]. (i) Following Albert’s terminology, we define the left and right multipli- cation maps La(b) := a · b and Ra(b) := b · a. (ii) We write Aλ(Xa) for the eigenspace of λ with respect to the trans- formation Xa, X ∈ {L, R}, i.e., Aλ(La)= {v ∈ A : a · v = λv}, and similarly for Aλ(Ra). Often we just write Aλ for Aλ(La), when a is understood. (iii) We denote: Aλ,δ(a) := Aλ(La) ∩ Aδ(Ra). An element in Aλ,δ(a) will be called a (λ, δ)-eigenvector of a, and (λ, δ) will be called its eigenvalue.

We just write Aλ,δ when the idempotent a is understood. Thus

Aλ,λ = {v ∈ A : a · v = λv = v · a}, cf. [HSS, §1.1, p. 263], and a ∈ A1,1. Of course Aλ,λ = Aλ when A is commutative. Lemma 2.3 gives identities in the operators La,Ra : A → A. We study idempotents in terms of the following notions. Definitions 1.2. (1) Idempotents a, b ∈ A are orthogonal if ab = ba = 0. (2) An idempotent a ∈ A is primitive if it cannot be written as the sum of nontrivial orthogonal idempotents. (3) An idempotent a ∈ A is left (resp. right) absolutely primitive if A1(La) = F a (resp. A1(Ra) = F a). The idempotent a is absolutely primitive if it is both left and right absolutely primitive. (Sometimes this is called “primitive” in the literature.) (4) An idempotent a ∈ A is flexible if LaRa = RaLa, and (xa)x = x(ax) for all x ∈ A. Lemma A (Lemmas 2.1 and 2.3). Suppose a ∈ A is a given flexible idem- potent. 2 2 (i) Ra − Ra = La − La. (ii) Ra(Ra + La − 1) = La(Ra + La − 1). (iii) If A is power-associative over a field of 6= 2, 3, then

(Xa − 1)Ya(La + Ra − 1) = 0, for X,Y ∈ {R,L}. Next, write ˚ A1/2(a)= {x ∈ A | ax + xa = x}. ˚ (We write A1/2, when a is understood.) Albert proved for A power-associative with char(F ) 6= 2, 3 that ˚ A = A1,1 ⊕ A0,0 ⊕ A1/2. This is reproved directly as Theorem 2.10(i). The following theorem of Albert then describes the appropriate fusion laws in the algebra A : FLEXIBLE IDEMPOTENTS IN NONASSOCIATIVE ALGEBRAS 3

Albert’s Fusion Theorem ([A, Theorem 3, p. 560, Theorem 5, p. 562]). Suppose that A is power-associative with char(F ) 6= 2, 3. Let a ∈ A be an idempotent.

(i) A0,0, A1,1 are subalgebras. Furthermore for A flexible, ˚ ˚ ˚ ˚ (ii) Aλ,λA1/2 ⊆ A1−λ,1−λ + A1/2, A1/2Aλ,λ ⊆ A1−λ,1−λ + A1/2, for λ = 0, 1. ˚ ˚ ˚ (iii) aA1/2, A1/2a ⊆ A1/2. Proposition 3.4 provides a useful decomposition result into eigenspaces. In §3 we drop power-associativity, and turn to a noncommutative version of axes, in preparation for the study of flexible axial algebras in [RS2]. The simplest nontrivial situation is for A1(La)= F a and for La to have exactly one eigenvalue ∈/ {0, 1}. (When the algebra A is not power-associative, the 1 third eigenvalue need not be 2 . This makes [T] quite surprising, giving a 1 condition for 2 to be an eigenvalue of an element of A after all.) Accordingly, we have: Definition 1.3. Let a ∈ A be an idempotent, and λ, δ∈ / {0, 1} in F . (1) a is a left axis of type λ if (a) a is left absolutely primitive, i.e., A1 = F a. (b) (La − λ)(La − 1)La = 0. (c) (cf. Proposition 3.4) There is a direct sum decomposition of A which is a Z2-grading: +-part −-part

A = zA0 }|⊕ A1{ ⊕ z}|{Aλ ,

recalling that Aλ means Aλ(La). We call this grading the left axial fusion rules. (2) A right axis of type λ is defined similarly. (3) a is an axis (2-sided) of type (λ, δ) if a is a left axis of type λ and a right axis of type δ and, in addition, LaRa = RaLa. Thus A1,1 = F a = A1; in particular A1,δ = Aλ,1 = 0. Hence − − −− A = A++ ⊕ A+ ⊕ A + ⊕ A , ′ ′ is Z2 × Z2 grading of A (multiplication (ǫ,ǫ )(ρ, ρ ) is defined in the obvious way for ǫ,ǫ′, ρ, ρ′ ∈ {+, −}), where ++ • A = A1 ⊕ A0,0, +− • A = A0,δ, −+ • A = Aλ,0, −− • A = Aλ,δ. (4) The axis a is of Jordan type (λ, δ) if Aλ,0 = A0,λ = 0. +− −+ In this case, A = A = 0, A0,0 = A0, and

A = A1 ⊕ A0 ⊕ Aλ,δ. 4 LOUIS ROWEN, YOAV SEGEV

(This definition is close to commutative, but there are natural noncom- mutative examples given in Examples 3.1 and 3.6.) Note that besides being noncommutative, Definition 1.3(1c) does not re- quire the condition that A0 is a subalgebra, contrary to the usual hypothesis in the commutative theory of axial algebras. But this condition does not seem to pertain to any of the proofs. If a is an axis of type (λ, δ), then by Definition 1.3(3), we can write x ∈ A as

x = αxa + x0,0 + x0,δ + xλ,0 + xλ,δ, xγ,ρ ∈ Aγ,ρ, which we call the (2-sided) decomposition of x with respect to a. When a is of Jordan type, we have the much simpler decomposition

x = αxa + x0 + xλ,δ with x0 ∈ A0, and xλ,δ ∈ Aλ,δ, which is both the left decomposition and right decomposition. Much of the axial theory (cf. [HRS]) can be generalized to this noncom- mutative setting, as exemplified in the following results Proposition B (Corollary 3.5).

(i) An axis a of Jordan type (λ, δ) is in C(A), iff either Aλ,δ = 0, or λ = δ. (ii) Let a 6= b be two axes with ab = ba and a of type (λ, δ). Then either ab =0= bλ,δ, or λ = δ. In particular, if a is of Jordan type, then either a ∈ C(A) or ab = 0.

Proposition C (Proposition 3.9). If A is an algebra generated by two axes a and b of Jordan type (whose types may be distinct!), then A = F a + F b + F ab. For flexible algebras we have Theorem D (Theorem 3.7). Suppose A is a flexible algebra. Then (1) Any axis a has some Jordan type (λ, δ), with either λ = δ, so that 2 a ∈ C(A), or λ + δ = 1 (λ 6= δ) and x = 0 for all x ∈ Aλ,δ. (2) Assume that both a and b are not in C(A). Then either ab = 0 or the subalgebra generated by a, b is one of the two examples as in Examples 3.6.

2. Basic properties of idempotents

In the following four lemmas (except in part (ii) of Lemma 2.3), one need not assume characteristic 6= 2.

2 2 Lemma 2.1. If a is a flexible idempotent, then La − La = Ra − Ra. FLEXIBLE IDEMPOTENTS IN NONASSOCIATIVE ALGEBRAS 5

Proof. We apply the definition twice: a + ax + (xa)a + xax = (a + xa)(a + x) = ((a + x)a)(a + x)= (a + x)(a(a + x)) = (a + x)(a + ax)= a + a(ax)+ xa + xax, so ax + (xa)a = a(ax)+ xa.  2 2 Lemma 2.2. Suppose a ∈ A is an idempotent satisfying La −La = Ra −Ra. (In particular this holds when a is a flexible idempotent, by Lemma 2.1.) If z is a (λ, δ)-eigenvector, then either δ = λ or δ = 1 − λ. Proof. (λ2 − λ)z = a(az) − az = (za)a − za = (δ2 − δ)z, implying λ2 − λ = δ2 − δ, or (λ − δ)(λ + δ − 1) = 0. 

Lemma 2.3. Suppose a ∈ A is a given flexible idempotent.

(i) Ra(Ra + La − 1) = La(Ra + La − 1). (ii) If A is power-associative over a field of characteristic 6= 2, 3, then

(Xa − 1)Ya(La + Ra − 1) = 0, for X,Y ∈ {R,L}. Proof. (i) By Lemma 2.1, 2 2 Ra(Ra + La − 1) = RaLa + Ra − Ra = RaLa + La − La = La(Ra + La − 1).

(ii) We show that (Ra − 1)Ra(La + Ra − 1) = 0, the rest follows from (i). By [A, Equation 12, p. 556], and applying Lemma 2.1, 2 2 3 2 0= Ra + LaRa + La − Ra − LaRa − La 2 2 3 2 (2.1) = Ra + LaRa + Ra − Ra − LaRa − Ra 2 3 2 = 2Ra − Ra + La(Ra − Ra) − Ra, so 2 3 2 2 (Ra − 1) Ra = Ra − 2Ra + Ra = La(Ra − Ra)= −LaRa(Ra − 1), so (Ra − 1)Ra(La + Ra − 1) = 0.  Theorem 2.4. (1) If a, a′ are orthogonal idempotents in A, then they are orthogonal in A(◦). (2) Any primitive idempotent of A(◦) is a primitive idempotent of A. (3) Suppose that A is flexible, and let a be an absolutely primitive idem- potent of A(◦). Let d ∈ A such that ad = d. Then (i) da = αa, and d2 = αd, for some α ∈ F. 1 (ii) If α = 0, then d is a (1, 0)-eigenvector of a, so a ◦ d = 2 d, and if α 6= 0, then (a − α−1d) is a (1, 0)-eigenvector of a, so −1 1 −1 a ◦ (a − α d)= 2 (a − α d). 6 LOUIS ROWEN, YOAV SEGEV

Proof. (1) This is obvious. (2) If a is a sum a′ + a′′ of orthogonal idempotents in A then a is a sum a′ + a′′ of orthogonal idempotents in A(◦), by (1). (3) (i) By Lemma 2.1, (da)a − da = a(ad) − ad = 0. Hence a ◦ da = 1 2 (ada + (da)a)= da, so by hypothesis da ∈ F a. Writing da = αa, we have d2 = d(ad) = (da)d = αa · d = αd. (ii) This is an easy computation. 

Note that Cases (i) and (ii) can arise in matrix algebras, respectively, with a = e11 and d = e12, or a = e11 + e12 and d = e11. Remark 2.5. Any algebra B generated by two elements a, v such that a2 = a, av = v, va = αa, and v2 = αv with α ∈ F, is flexible and satisfies A1(La)= B and A1(Ra)= F a. 2.1. Semisimple idempotents.

Definition 2.6. An idempotent a is left semisimple of degree t if

t Y(La − λi) = 0 i=1 for suitable distinct eigenvalues λ1, . . . , λt. The idempotent is semisimple if it is both left and right semisimple.

Proposition 2.7. Suppose a is a left semisimple idempotent of degree t. For y ∈ A, write y = P yj where yj is the left λj-eigenvector in the left eigen- t−1 i vector decomposition of y, and define the vector space Va(y)= Pi=0 F L ay. Then t Va(y)= ⊕j=1Fyj.

k k Proof. By induction, La(y)= P λj yj for 0 ≤ k ≤ t − 1. Thus t Va(y) ⊆⊕j=1Fyj.

But on the other hand, we have a system of t equations in the yj with k coefficients (λj ), the Vandermonde matrix, which is nonsingular, so there is a unique solution for the yj in Va(y). 

This space Va(b) plays a key role for an idempotent b. Here is another useful general result.

Proposition 2.8. If a is a semisimple idempotent of a flexible algebra A, then A decomposes as ⊕λAλ,λ ⊕ Aλ,1−λ, summed over all left eigenvalues λ of a.

Proof. These are the only possibilities, in view of Lemma 2.2.  FLEXIBLE IDEMPOTENTS IN NONASSOCIATIVE ALGEBRAS 7

2.2. Idempotents in power-associative algebras. Albert [A] proves some amazing results. [A, Equation (20), p. 568] im- plies that if char(F ) 6= 2, 3, then every idempotent in a commutative power- is an axis (see the Introduction and Definition 1.3). Fur- thermore, the eigenspaces are independent by [A, Equation (23), p. 569]. [A, Theorem I.2, p. 559] computes the Albert fusion rules with respect to a given idempotent a:

(i) A1A1 ⊆ A1 and A0A0 ⊆ A0; (ii) A0A1 = A1A0 = {0}; (iii) A0A1/2 ⊆ A1/2 + A1; (iv) A1A1/2 ⊆ A1/2 + A0; 2 (v) A1/2 ⊆ A0 + A1. We stress the role here of power-associativity. In Theorems I.3, I.5 ([A, pp. 560, 562]), Albert gets the same results, ex- cept for (v), for (noncommutative) power-associative algebras with char(F ) 6= 2, 3, replacing the eigenspaces Aλ, with the ◦-eigenspaces:

A˚λ := {x ∈ A | ax + xa = 2λx}.

(These are Aa(λ) in the notation of Albert, [A, Equation (27), p. 560].) He furthermore shows that A˚λ = Aλ,λ = Aλ, and that these are subrings, for λ ∈ {0, 1}. For Jordan algebras cf. [A, Equation (26), p. 559] Albert improves his fusion rules:

(i) A1A1 ⊆ A1 and A0A0 ⊆ A0; (ii) A1A0 = {0}; (iii) (A0 + A1)A1/2 ⊆ A1/2; 2 (iv) A1/2 ⊆ A0 + A1, but [A, Equation (26), p. 559, l.-2] has a counterexample to (iii) in a com- mutative power-associative algebra. When (iii) does hold, Albert calls the algebra A stable. Many of the above results can be extended to the noncommutative situ- ation. Remark 2.9. As we mentioned, in an easy argument in the first three lines of the proof of [A, Theorem 3, p. 560] Albert proves that A˚λ = Aλ,λ for λ ∈ {0, 1}. This is used in the following result of Albert on power associative rings ([A, (22), p. 559] and [A, Thm. 3, p. 562]). Theorem 2.10 ([A]). Suppose that A is power-associative, and that char(F ) 6= 2, 3. Let a ∈ A be an idempotent. Then ˚ (i) A = A1,1 ⊕ A0,0 ⊕ A1/2. (ii) We have

A1 = Xa(La + Ra − 1)A; A0 = (Xa − 1)(La + Ra − 1)A; ˚ A1/2 = (Ra + La)(Ra + La − 2)A, X ∈ {R,L}. 8 LOUIS ROWEN, YOAV SEGEV

Here is an alternate proof that is rather conceptual, working directly in A. Proof. (i) Assume first that A is commutative. Then A is flexible (see equa- tion (1.1)), so we may use Lemma 2.3. Let x1 = Rax and x2 = (1 − Ra)x, so that x = x1 + x2. By Lemma 2.3(ii) we have

(2.2) (Ra − 1)(Ra + La − 1)(x1)=0=(La − 1)(Ra + La − 1)(x1). and

(2.3) Ra(Ra + La − 1)(x2)=0= La(Ra + La − 1)(x2). Let ′ ′′ x1 = (Ra + La − 1)(x1), x1 = (Ra + La − 2)(x1), ′ ′′ x2 = (Ra + La − 1)(x2), x2 = (Ra + La)(x2), Of course ′ ′′ ′′ ′ x = x1 − x1 + x2 − x2. By Equations (2.2) and (2.3), ′ ′ ′ ′ ′ Ra(x1)= x1 = La(x1),Ra(x2)=0= La(x2). So ′ ′ x1 ∈ A1, and x2 ∈ A0. ′′ ′′ ˚ Also by Equation (2.2), (Ra + La − 1)(x1 ) = 0, that is x1 ∈ A1/2, and by ′′ ′′ ˚ Equation (2.3), (Ra + La − 1)(x2 ) = 0, that is x2 ∈ A1/2. For the general case where A is not commutative, A˚ is commutative and ˚ ˚ ˚ ˚ power associative; hence A = A1 ⊕ A0 ⊕ A1/2. However, Aλ = Aλ,λ, for λ ∈ {0, 1}, by Remark 2.9. (ii) follows from (i). Indeed ˚ Xa(La + Ra − 1)A0 =0= Xa(La + Ra − 1)A1/2, and Xa(La+Ra−1)(x)= x, for x ∈ A1, and similarly for (Xa−1)(La+Ra−1) and (Xa + La)(Ra + La − 2).  ˚ Note that if A is commutative then A1/2 = A1/2, so Theorem 2.10 gener- alizes the commutative description of idempotents. 2.2.1. Application to noncommutative Jordan algebras.

Definition 2.11. ([Sch]) An algebra A is noncommutative Jordan if A is flexible and satisfies the identity (2.4) (x2w)x = x2(wx). Schafer [Sch, p. 473] showed that any noncommutative with 1 of characteristic 6= 2 satisfying (2.4) is flexible and power-associative. Hence, Albert’s Fusion Theorem applies to noncommutative Jordan algebras with 1. FLEXIBLE IDEMPOTENTS IN NONASSOCIATIVE ALGEBRAS 9 3. Basic properties of axes

In this section we generalize certain notions from the theory of commu- tative axial algebras in [HRS], to the noncommutative setting. We start with some examples of axes that are not flexible. Examples 3.1. (i) Let A = F a + F b + F c with multiplication given by a2 = a, b2 = a = c2, ab = λb, ba = 0, ac = 0, ca = δc, bc = cb = 0.

Then a(ba)=0=(ab)a, and a(ca)=0=(ac)a, so LaRa = RaLa.

(La − λ)b = λ(b − b)=0= Lac,

implying (La − λ)(La − 1)La = 0. Likewise for Ra. Hence a is an axis of type (λ, δ) which is not Jordan. It is the only nonzero axis in A. (ii) Let 0 6= λ, δ ∈ F, with λ + δ = 1. Let ′ A = F a ⊕ Fx ⊕ Fy ⊕ Fy ⊕ F z, with multiplication defined by: ′ ′ ′ a2 = a, ax = xa = 0, ay = λy, ya = 0, ay = 0,y a = δy , az = λz, za = δz.

′ ′ x2 = xy = yx = xy = y x = xz = zx = 0. ′ ′ ′ y2 = −x, yy =0= y y, yz = λy ,zy = 0. ′ ′ ′ (y )2 = x, y z = δy,zy = 0. z2 = 0. So ′ A1,1 = F a, A0,0 = Fx, Aλ,0 = Fy, A0,δ = Fy , Aλ,δ = F z. It is easy to check that (av)a = a(va), for v ∈ {a, x, y, y′, z}, and a is an axis of type (λ, δ). Of course a is not of Jordan type. Also, a is not a flexible axis since (ya)y = 0, while y(ay)= λy2 = −λx. Let b = a + x + y + y′ + z. Then ′ ′ ′ b2 =a2 + y2 + (y )2 + ay + y a + az + za + yz + y z ′ = a − x + x + (λ + δ)y + (λ + δ)y + (λ + δ)z = b, (We conjecture that by solving some more equations b could also be made into an axis.) Note that A has dimension 5. Since a, b, ab, ba, and aba are independent, they span A.

Remark 3.2. Write y = y0 + αya + yλ, for yρ ∈ Aρ, ρ ∈ {0, λ}, and α ∈ F . In this notation, y0 = y0,0 + y0,δ, and yλ = yλ,0 + yλ,δ. When a has Jordan type (λ, δ), then y0 = y0,0 and yλ = yλ,δ. 10 LOUIS ROWEN, YOAV SEGEV

Lemma 3.3. Suppose a is an axis of type (λ, δ).

(1) a(A0 + F a)= F a. (2) ay = αya + λyλ, implying 1 y = (ay − α a). λ λ y (3) a(ay)= αy(1 − λ)a + λay. 1 (4) y0 = y − λ (ay − (λ + 1)αya). (5) (ay)a = αya + λδyλ,δ, implying 1 y = ((ay)a − α a). λ,δ λδ y 2 Proof. (1) If x0 ∈ A0 then ax0 = 0, so a(A0 + A1)=0+ aF a = F a = F a. (2) Apply La to Remark 3.2. 2 (3) a(ay)= αya + λ yλ = αya + λ(ay − αya), yielding (3). 1 (4) y0 = y − yλ − αya = y − λ (ay − (λ + 1)αya). (5) Apply Ra to (2).  Proposition 3.4. (i) If a is a left axis of type λ then the left decomposition

x = x1 + x0 + xλ of x ∈ A with respect to a, is given by: 1 • x1 = λ−1 (a(a − λ))x = αxa, αx ∈ F, 1 • x0 = − λ ((a − λ)(a − 1))x ∈ A0(La), 1 • xλ = − λ(λ−1) (a(a − 1))x ∈ Aλ(La). (ii) If a is a right axis of type δ then we have an analogous further right decomposition xρ = xρ,1 + xρ,0 + xρ,δ of xρ ∈ Aρ, (for ρ ∈ {0, 1, λ} with respect to a, given by: 1 • xρ,1 = δ−1 xρ(a(a − δ)) = αxρ a, αxρ ∈ F, 1 • xρ,0 = − δ xρ((a − δ)(a − 1)) ∈ A0(Ra), 1 • xρ,δ = − δ(δ−1) xρ(a(a − 1)) ∈ Aδ(Ra). (iii) An idempotent b cannot be a left eigenvector of a left axis a 6= b unless ab = 0. (iv) If a is an axis of type (λ, δ), then A decomposes into a direct sum ++-part +−-part −+-part −−-part

A = zA1,1 }|⊕ A0,0{ ⊕ z}|{A0,δ ⊕ zA}|λ,0{ ⊕ Az}|λ,δ{ ,

and this is a Z2 × Z2-grading of A. In particular A0,1 = Aλ,1 = A1,0 = A1,δ = 0. (v) If a is an axis of Jordan type (λ, δ) then +-part −-part

A = zA1,1 }|⊕ A0,0{ ⊕ zA}|λ,δ{ ,

is a Z2-grading of A. FLEXIBLE IDEMPOTENTS IN NONASSOCIATIVE ALGEBRAS 11

Proof. Note that

1 1 1 (3.1) λ−1 (La(La − λ)) − λ ((La − λ)(La − 1)) − λ(λ−1) (La(La − 1)) = 1, so their sum applied to x is x, and each term is in the appropriate left eigenspace. (ii) As in (i), using equation (3.1), with Ra in place of La, noting that La and Ra commute. (iii) If the idempotent b 6= a is a left eigenvector, then, since a is left absolutely primitive, its eigenvalue must be 0 or λ. But if b ∈ Aλ(La), then, 2 2 by Definition 1.3(1c), b ∈ A0(La)+ A1(La), but b = b, a contradiction. (iv) Combining (i) and (ii) together with the left Z2 and the right Z2 grading given in Definition 1.3((1)& (2)), shows that:

++-part +−-part −+-part −−-part

A = Az 1,1 ⊕ A1,0 }|⊕ A0,1 ⊕ A0,{0 ⊕ zA1,δ }|⊕ A0,δ{ ⊕ zAλ,1 }|⊕ Aλ,0{ ⊕ zA}|λ,δ{ , is a Z2 × Z2 grading of A. We have αxa = x1,1 + x1,0 + x1,δ, so x1,1 = αxa and x1,0 = x1,δ = 0. Working from the other side yields x0,1 = xλ,1 = 0. (v) The other components vanish, by the definition of Jordan type.  Corollary 3.5.

(i) An axis a is in the center of A, iff it is of Jordan type with Aλ,δ = 0. (ii) An axis a of Jordan type (λ, δ) is in C(A), iff either Aλ,δ = 0 or λ = δ. (iii) Let a 6= b be two axes with ab = ba and a of type (λ, δ). Then either ab = 0, or λ = δ. In particular, if a is of Jordan type, then either a ∈ C(A), or ab = 0.

Proof. (i) (⇒) For any x ∈ A,

2 λ xλ + αxa = a(ax)= ax = λxλ + αxa, implying xλ = 0 since λ 6= 0, 1. Hence xλ,0 =0= xλ,δ, and by symmetry x0,δ = 0. (⇐) x = αxa + x0,0 for any x in A, implying ax = xa = αxa and also xy = x0,0y0,0 + αxαya, and thus a(xy) = (ax)y = αxay = αxαya for all x,y. (ii) Suppose that a ∈ C(A) and that Aλ,δ 6= 0. Let 0 6= xλ,δ ∈ Aλ,δ. Then λx = ax = xa = δx, so λ = δ. Conversely, for x ∈ A, decompose x according to a : x = αa + x0,0 + xλ,λ, where α ∈ F and xλ,λ may be 0. Then ax = αa + λxλ,λ = xa, so a ∈ C(A). (iii) Let a be of type (λ, δ) and let b = αba + b0,0 + bλ,0 + b0,δ + bλ,δ be the decomposition of b with respect to a. Then αba + λbλ,0 + λbλ,δ = ab = ba = αba + δb0,δ + δbλ,δ implying bλ,0 = b0,δ = 0. If bλ,δ 6= 0, then λ = δ, and we are done. Hence bλ,δ = 0, in which case ab ∈ F a. But then a is an eigenvector of b, so by Proposition 3.4(iii), ab = 0.  12 LOUIS ROWEN, YOAV SEGEV

3.1. Axes in flexible algebras. Next we analyze the situation in flexible algebras. Here are two examples of flexible algebras that are not commutative. Examples 3.6. Let λ, δ ∈ F such that λ, δ∈ / {0, 1}, λ + δ = 1, and λ 6= δ. (1) Let A be the 2-dimensional algebra F a + F b with multiplication defined by a2 = a, b2 = b, ab = δa + λb ba = λa + δb. Then (ab)a = δa+λba = δa+λ2a+λδb and a(ba)= λa+δab = λa+δ2a+λδb. Note however that δ + λ2 = λ + δ2. So (ab)a = (ab)a. By symmetry (ba)b = 2 2 b(ab). Next we show that Ra − Ra = La − La. We have a(ab) − ab = a(δa + λb) − δa − λb = λab − λb = λ(δa + λb) − λb = λδa + λ2b − λb. and (ba)a − ba = (λa + δb)a − λa − δb = δba − δb = δ(λa + δb) − δb = δλa + δ2b − δb. Since λ2 − λ = δ2 − δ, we are done. Next ((αa + βb)a)(αa + βb) = (αa + βba)(αa + βb)= α2a + αβab + βα(ba)a + β2(ba)b, and ((αa + βb)(a(αa + βb)) = (αa + βb)(αa + βab)= α2a + αβa(ab)+ βαba + β2b(ab). Since (ba)a − ba = a(ab) − ab, and since (ab)a = a(ba), we see that (xa)x = x(ax), for all x ∈ A. By symmetry (xb)x = x(bx), for all x ∈ A, so A is flexible. We have a(a − b)= a − ab = a − (δa + λb)= λ(a − b) and b(a − b)= ba − b = λa + δb − b = λ(a − b). Similarly (a − b)a = δ(a − b) = (a − b)b. (Indeed (a − b)2 = 0.) We thus 2 have Aλ,δ(a) = Aλ,δ(b) = F (a − b), with (F (a − b)) = 0, and of course A0(a) = A0(b) = 0. Thus the fusion laws hold for both a and b, and both are axes of Jordan type (λ, δ). Note that the idempotents in A have the form αa + (1 − α)b, α ∈ F. (2) Let A = F a + F b + F x, with multiplication defined by a2 = a, b2 = b, x2 = 0, and ab = ax = xb = λx, ba = xa = bx = δx. FLEXIBLE IDEMPOTENTS IN NONASSOCIATIVE ALGEBRAS 13

So Fx = Aλ,δ, A0,0 = F (b − x). In checking the fusion rules for a, we have (b−x)x = bx ∈ Fx and x2 = 0, and also (b − x)2 = b2 − bx − xb + x2 = b − (λ + δ)x = b − x. Likewise for b. Thus a and b are axes of Jordan type (λ, δ) and (δ, λ) respectively. Let us check the eigenvalues of the idempotent b − x. a(b − x)=0= (b − x)a. (b − x)x = δx, and x(b − x)= λx. Thus b − x also is a Jordan axis, of type (δ, λ). Note that the idempotents in A all have the form αa + βb + ξx, with α + β = 1. To show that A is flexible, we compute that ((αa + βb + ξx)a)(αa + βb + ξx) = (αa + ξx)(αa + βb + ξx) η = βδ + ξδ = α2a + αβλx + αξλx + η(αδ + βλ)x = α2a + αβλx + αξλx + (αβδ2 + β2λδ + αξδ2 + ξδβλ)x and (αa + βb + ξx)(a(αa + βb + ξx)) = (αa + βb + ξx)(αa + γx) γ = βλ + ξλ = α2a + αβδx + αξδx + γ(αλ + βδ)x = α2a + αβδx + αξδx + (αβλ2 + β2λδ + ξαλ2 + ξλβδ)x We must show that αβλ + αξλ + αβδ2 + αξδ2 = αβδ + αξδ + αβλ2 + ξαλ2. But this follows from the fact that λ2 − λ = δ2 − δ. By symmetry (yb)y = y(by), for all y ∈ A. It is easy to check that (yx)y = y(xy), for all y ∈ A, and we see that A is flexible. Theorem 3.7. Suppose A is a flexible algebra. Then (1) Any axis a has some Jordan type (λ, δ) and is either in C(A) (with 1 2 λ = δ), or δ = 1 − λ, λ 6= 2 , and x = 0 for all x ∈ Aλ,δ. (2) For axes a, b ∈ A, write

b = αba + b0 + bλ,δ and a = αab + a0 + aλ′,δ′ ,

where b0 ∈ A0(a), bλ,δ ∈ Aλ,δ(a), a0 ∈ A0(b), aλ′,δ′ ∈ Aλ′,δ′ (b). Assume that λ 6= δ, and λ′ 6= δ′ (so that both a and b are not in C(A), and δ = 1 − λ and δ′ = 1 − λ′). 2 2 (I) (ab) = αb(ab), and (ba) = αb(ba). In particular, if ab 6= 0, then αa = αb. (II) Suppose that bλ,δ 6= 0, so that ab 6= 0, and hence αa = αb. Write α := αb, and let ha, bi be the subalgebra generated by a, b. Then 14 LOUIS ROWEN, YOAV SEGEV

(i) b0bλ,δ = δ(1 − α)bλ,δ, and bλ,δb0 = λ(1 − α)bλ,δ. (ii) bbλ,δ = (α(λ − δ)+ δ)bλ,δ and bλ,δb = (α(δ − λ)+ λ)bλ,δ. In particular aλ′,δ′ is a scalar multiple of bλ,δ. (iii) ha, bi is spanned by a, b, bλ,δ. (iv) If α 6= 0, then α = 1, and ha, bi = F a + F b is isomorphic to the algebra in Example 3.6(i). Hence λ′ = λ. (v) If α = 0, then ha, bi is isomorphic to the algebra in Ex- ample 3.6(ii). Hence λ′ = 1 − λ. (III) Either ab = ba = 0 or ha, bi is isomorphic to one of the algebras in Examples 3.6. Proof. (1) The first assertion is by Proposition 2.8. Suppose a is an axis of 2 2 Jordan type (λ, δ). Then δx = (xa)x = x(ax) = λx , for x ∈ Aλ,δ. Hence 2 λ = δ unless x = 0 for all x ∈ Aλ,δ. (2)(I) We have −1 ab = αba + λbλ,δ =⇒ bλ,δ = λ (ab − αba). 2 Hence, since abλ,δ + bλ,δa = bλ,δ, and bλ,δ = 0, we get 2 2 2 (ab) = (αba + λbλ,δ) = αb a + αb(ab − αba)= αb(ab). 2 The proof that (ba) = αb(ba) is similar. (2)(II) (i) We have 2 2 b = b = (αba + b0 + bλ,δ) 2 2 = αb a + b0 + αb(λ + δ)bλ,δ + b0bλ,δ + bλ,δb0.

Hence, by the Z2 grading of A, we have 2 2 (3.2) b0 = (αb − αb )a + b0, and

(3.3) (1 − αb)bλ,δ = b0bλ,δ + bλ,δb0. Furthermore, linearizing the flexible axiom yields (xy)z + (zy)x = x(yz)+ z(yx).

Taking b0 for x and a for y yields xy = yx = 0, implying (za)b0 = b0(az), and thus, for z = bλ,δ we have

(3.4) δbλ,δb0 = λb0bλ,δ. Combining equations (3.3) and (3.4) and the fact that λ + δ = 1, we get (i).

(ii) This is an easy computation using (i), λ + δ = 1, and the fact that 2 bλ,δ = 0.

(iii) Set B := F a + F b + F bλ,δ. Clearly ab, abλ,δ, ba, bλ,δa ∈ B. By (ii), also 2 bbλ,δ, bλ,δb ∈ B, and bλ,δ = 0. FLEXIBLE IDEMPOTENTS IN NONASSOCIATIVE ALGEBRAS 15

(iv) Suppose α 6= 0. By (ii) we have

ab = αb + γbλ,δ, γ ∈ F.

Hence

0= ab − ab = α(a − b)+ ρbλ,δ, ρ ∈ F.

This shows that bλ,δ is a multiple of a − b. By (iii), ha, bi = F a + F b. In particular ha, bi0(b) = 0, so b = αa + bλ,δ. Since bλ,δ is a multiple of a − b, we must have α = 1. It is now easy to check that (iv) holds.

(v) Suppose that α = 0. Then, by equation (3.2), b0 is an idempotent, and by (i),

b0bλ,δ = δbλ,δ and bλ,δb0 = λbλ,δ.

But now b0 is an idempotent and b = b0 + bλ,δ, where bλ,δ ∈ Aδ,λ(b0). Also bb0 6= b0b. By (iv) the subalgebra generated by b and b0 is as in Example 3.6(i). Hence, since b0 is of type (δ, λ), b is also of type (δ, λ). Now ab = λaδ,λ, but, since α = 0, also ab = λbλ,δ. Hence aδ,λ = bλ,δ. We now see that ha, bi is the algebra as in Example 3.6(ii), with x = bλ,δ.

(III) Assume that ab 6= 0. If bλ,δ = aλ′,δ′ = 0, then ab = αba = ba = αab. this is possible iff αa = αb = 0, but then ab = 0. Hence we may assume without loss that bλ,δ 6= 0, and (III) holds by (II).  Theorem 3.8. Suppose a is a left axis, and b is a right axis. Then (i) a(ax) ∈ F a + F ax, for all x ∈ A. (ii) (xb)b ∈ F b + F xb, for all x ∈ A. (iii) Let V = F a + F b + F ab + F ba. Then (ab)(ab) − a(bab) ∈ V. (iv) Suppose that a and b are axes (whose types may be distinct!), and that A is generated by a and b. (a) If V (of (iii)) contains aba, bab, (ab)(ba), and (ba)(ab), then V = A. (b) Let ′ V = F a + F b + F ab + F ba + F aba + F bab.

Then aV ′, bV ′,V ′a, V ′b ⊆ V ′.

Proof. (i) , (ii) Special cases of Lemma 3.3.

(iii) We write y =∼ z if y − z ∈ V. Now (ab − λb) ∈ F a + A0(La), so

bλ(ab − λb) ∈ Aλ. 16 LOUIS ROWEN, YOAV SEGEV

Note that by (ii), (ab)b ∈ V. Note also that applying La to (i) implies a(a(ab)) ∈ V. Hence ∼ ∼ (ab)(ab) = ab(ab − λb)= λbλ(ab − λb)+ αba(ab − λb) = λbλ(ab − λb)

= a(bλ(ab − λb)) = a((b − αba − b0)(ab − λb))

= a((b − αba)(ab)) − λa((b − αba)b) − a(b0(ab − λb))

= a(bab) − αba(a(ab)) − λab + λαba(ab) − a(b0(ab − λb)) =∼ a(bab), in view of (i), since a(b0(ab − λb)) ∈ a(A0 + F a) ⊆ F a. (iv(a)) We check that V is closed under products. By (i) and (ii) and since aba, bab ∈ V, we see that V is closed under La,Ra,Lb and Rb. By (iii), (ab)(ab) =∼ a(bab) ∈ aV ⊆ V . Reversing the roles of a and b shows that (ba)(ba) ∈ V . We are given that (ab)(ba) ∈ V and (ba)(ab) ∈ V. (iv(b)) By symmetry (both with respect to a, b and to working on the left or on the right), it is enough to show that aV ′ ⊆ V ′. By (i), a(aba) = a(a(ba)) ∈ F a + F aba ⊆ V ′. It remains to check that a(bab) ∈ V ′. By (iii) and (i), ′ a(ab)(ab) − a(bab) ∈ V . By Lemma 3.3(3), a(a(bab))−λa(bab) ∈ V ′ implying a((ab)(ab))−λa(bab) ∈ V ′. Let the type of a be (λ, δ). We work with the decomposition of b with ′ ′ respect to a. Now aba = αba + λδbλ,δ, so bλ,δ ∈ V . Hence bλ,0 ∈ V , because ′ ab = αba+λbλ,0 +λbλ,δ ∈ V . by the fusion rules of Proposition 3.4(iv), that a((ab)(ab)) is an F -linear combination of a, bλ,0 and bλ,δ. Hence a((ab)(ab)) ∈ V ′, so a(bab) ∈ V ′.  In [RS2, Theorem 1.10] we shall prove that the conclusion of (iv(b)) im- plies A = F a + F b + F ab + F ba + F aba + F bab, by means of an extra technique of “Miyamoto involutions.” This result can be improved in cer- tain situations: Corollary 3.9. Suppose A is an algebra generated by two axes a and b (whose types may be distinct!), and let V = F a + F b + F ab. (i) If ba ∈ V , then V = A. (ii) In particular, (i) is the case when a or b are axes of Jordan type.

Proof. (i) By Theorem 3.8(i), V is closed under La. In particular aba ∈ V, so V is closed under Ra. Similarly V is closed under Rb by Theorem 3.8(ii), so bab ∈ V, and then V is closed under Lb. It remains to show that (ab)(ab) ∈ V, which follows from Theorem 3.8(iii). (ii) (For a of Jordan type) ba = αba + δbλ,δ, and bλ,δ ∈ V by Theo- rem 3.8(i), so ba ∈ V.  FLEXIBLE IDEMPOTENTS IN NONASSOCIATIVE ALGEBRAS 17

We cannot improve (ii) directly to arbitrary axes. However, we have a positive result, which will be developed further in [RS2]. Proposition 3.10. Let a be an axis of type (λ, δ). For x ∈ A, write x = αxa + x0,0 + xλ,0 + x0,δ + xλ,δ as in (1), and define the vector space

Va(x)= F a ⊕ Fx ⊕ F ax ⊕ F xa ⊕ F axa. Then Va(x)= F a ⊕ Fx0,0 ⊕ Fxλ,0 ⊕ Fx0,δ ⊕ Fxλ,δ. Proof. Applying a special case of Proposition 2.7 both on the left and right, 2 i j X FLaRax = F a ⊕ Fx0,0 ⊕ Fxλ,0 ⊕ Fx0,δ ⊕ Fxλ,δ. i,j=0 2 2 We conclude with Lemma 3.3(3), which lets us replace La by La, and Ra by Ra (when we add F a). 

References

[A] A.A. Albert, Power-associative rings, Trans. Amer. Math. Soc. 64 (1948), 552– 593. [DPSC] Tom De Medts, S.F. Peacock, D. Shpectorov, and M. Van Couwenberghe, De- composition algebras and axial algebras, arXiv:1905.03481v3 [math.RA], 25 Aug 2020. [HRS] J. I. Hall, F. Rehren, S. Shpectorov, Primitive axial algebras of Jordan type, J. Algebra 437 (2015), 79–11. [HSS] J. I. Hall, Y. Segev, S. Shpectorov, Miyamoto involutions in axial algebras of Jordan type half, Israel Journal of Mathematics 223 (2018), 261–308. [RS1] L. Rowen and Y. Segev, Associative and Jordan algebras generated by two idem- potents, Algebras and Representation Theory 20(6) (2017), 1495–1504. [RS2] L. Rowen and Y. Segev, Finitely generated axial algebras, preprint (2021). [Sch] R.D. Schafer, Noncommutative Jordan algebras of characteristic 0, Proc. Amer. Math. Soc. 6 (1955), 472–475. [T] V.G. Tkachev, Habil, The universality of one half in commutative nonassociative algebras with identities, J. Algebra, to appear.

Louis Rowen, Department of Mathematics, Bar-Ilan University, Ramat Gan, Israel Email address: [email protected]

Yoav Segev, Department of Mathematics, Ben-Gurion University, Beer- Sheva 84105, Israel Email address: [email protected]