Poisson kernel and blow-up of the second derivatives near the boundary for Stokes equations with Navier boundary condition
Abstract.
We derive the explicit Poisson kernel of Stokes equations in the half space with nonhomogeneous Navier boundary condition (BC) for both infinite and finite slip length. By using this kernel, for any , we construct a finite energy solution of Stokes equations with Navier BC in the half space, with bounded velocity and velocity gradient, but having unbounded second derivatives in locally near the boundary. While the Caccioppoli type inequality of Stokes equations with Navier BC is true for the first derivatives of velocity, which is proved by us in [CPAA 2023], this example shows that the corresponding inequality for the second derivatives of the velocity is not true. Moreover, we give an alternative proof of the blow-up using a shear flow example, which is simple and is the solution of both Stokes and Navier–Stokes equations.
Key words: Navier boundary condition, Stokes system, Poisson kernel, Navier–Stokes equations, local regularity, boundary blowup
AMS Subject Classification (2000): 35Q30
1. Introduction
We consider the following Stokes system in , ,
(1.1) |
with nonhomogeneous Navier boundary condition (Navier BC) on
(1.2) |
for . The system can be also considered for with an initial condition at . Here is the velocity field, is the pressure, is the friction coefficient, and is the boundary value. For , (1.2) is reduced to Lions boundary condition. When , its inverse has the unit of length and is called the slip length. We refer to our previous paper [4] and the reference therein for more detailed introduction of the physical meaning, and the historical study on the Stokes and Navier–Stokes equations under Navier BC.
In this paper, we continue to study the local regularity of the solution of the Stokes equations (1.1) near boundary. We recall some developments along this line. For the Stokes system in the half space with Dirichlet BC, Seregin [15, Lemma 1.1] showed spatial smoothing (, ) if one assumes , where denotes parabolic cylinders (see Section 2 for definition). However, without any assumption on the pressure, the smoothing in spatial variables fails due to non-local effect of the pressure. The first counter example constructed by K. Kang [7] shows that there exists a bounded weak solution of the Stokes equations whose normal derivative is unbounded near boundary,
Seregin-Šverák [13] found a simplified example in the form of a shear flow such that its gradient is unbounded near boundary, although the velocity field is locally bounded. Recently, Chang-Kang [3, Theorem 1.1] proved that for any , a bounded very weak solution (depending on ) can be constructed whose derivatives are unbounded in ,
(1.3) |
See [2, 8, 9, 14] for related study along this line and [6] for partial regularity results on the boundary with Dirichlet BC.
For Stokes system (1.1) in the half space with Navier BC, we proved in [4, Theorem 1.1] the following Caccioppoli’s inequality,
(1.4) |
It is a very important distinction in comparison to the Dirichlet BC where Caccioppoli’s inequality fails near boundary by (1.3). Moreover, we also proved in [4, Theorem 1.3] that , for bounded velocity , if the friction coefficient ,
(1.5) |
Similar study was initiated by Dong-Kim-Phan [5], where boundary second derivative estimates for generalized Stokes system with VMO coefficients under Lions BC were proved. The main motivation of this paper is to study the regularity criteria of higher order derivatives of near the boundary when assuming non-zero friction coefficient. It will be proved that can not be replaced by in inequality (1.4).
At first, we derive the explicit Poisson kernel of Stokes system with Navier BC, which allows us to express and in terms of .
Theorem 1.1 (Poisson kernel).
By we mean functions with compact support.
We refer to [4, Definition 2.2] for the motivation of the definition of very weak solutions. Above we have added nonhomogeneous boundary data in the definition. Next we use the Poisson kernel to construct a finite energy solution of Stokes equations, which has bounded velocity and velocity gradient, while its second derivatives blow up in norm.
Theorem 1.2 (Unbounded second derivative in ).
Let , , . There exists a very weak solution of Stokes equations (1.1) in with homogeneous (i.e. ) Navier boundary condition (1.2) on such that
(1.8) |
Moreover, it is the restriction of a global solution of (1.1) in with finite global energy
(1.9) |
and this global solution satisfies Navier boundary condition (1.2) on with bounded and compactly supported boundary data .
Remark 1.3.
(i) This theorem under Navier BC is parallel to (1.3) under zero BC. It tells us that the similar inequality of (1.4) with replaced by is not true when . By (1.5), the blow-up of does not happen if .
(ii) One of the key ingredients in our proof is the inequalities (2.11) of Lemma 2.8 for the norm in the Besov space . The first inequality of (2.11) is proved by Chang-Kang [3] using interpolation theorems in anisotropic Sobolev spaces. We give a simple, alternative proof of full (2.11) in this paper.
(iii) The blow-up of will not happen if we assume regularity of the pressure. Actually in [4, Theorem 1.2], we proved that for ,
(1.10) |
(iv) An example can be constructed to make while and are bounded. See Remark 6.3.
Inspired by Seregin-Šverák [13], where they construct a shear flow in with unbounded derivatives of velocity near the boundary under Dirichlet BC, we will give another example which has the same kind of blow-up as (1.8). The advantage of this example is that it is simple and is the solution of both Stokes and Navier–Stokes equations. It should be pointed out that it is a shear flow and has no spatial decay. In particular, it has infinite global energy.
Theorem 1.4 (Shear flow example).
In fact, the same proof can be used to give a second example of (1.3) under zero BC. See Remark 6.2.
The rest of this paper is organized as follows: We introduce notations and some preliminary results in Section 2. We derive the expression of the Poisson kernel in Section 3. We give the estimate of the kernel and prove Theorem 1.1 in Section 4. We introduce the example for Theorem 1.2 in Section 5. We give a proof of Theorem 1.4 in Section 6.
2. Notations and preliminaries
For , denote as the horizontal variable. Let , , and . Denote the heat kernel
and the fundamental solution of
Here denotes the volume of the unit ball in . Notice that for ,
(2.1) |
Here we add the derivative to avoid singularity at when . Let be the Heaviside function. Denote if there is a constant such that , and depends on some variables which are clear in the context. Let be the convolution in terms of . We use the following definition of Fourier transform
and
(2.2) |
The following several lemmas (Lemma 2.1–Lemma 2.5) will be used in calculating the inverse Fourier transform to get the Poisson kernel. They are not new and are widely used for the derivation of Green tensor of the Stokes equations [9, 11]. We give their proofs here for the convenience of the readers.
Lemma 2.1.
Suppose . For we have
For , we have
Proof.
It is well known that the Fourier transform of the fundamental solution of is
Hence,
(2.3) |
We can use Residue Theorem to calculate the integral in (2.3). Since , is bounded when is in the upper half of the complex plane. There is only one singular point in the upper half of the complex plane, so
For , we calculate them directly:
Remark 2.2.
By differentiating the equation in Lemma 2.1 with respect to , we have
Corollary 2.3.
For and , we have
Lemma 2.4.
For and , we have
Note that in our definition, no matter or .
Proof.
It is well known that the Fourier transform of heat kernel is
Hence,
(2.6) |
We can use Residue Theorem to calculate the integral in (2.6). Since , is bounded when is in the upper half of the complex plane. There is only one singular point in the upper half of the complex plane, so
By differentiating the equation in Lemma 2.4 with respect to , we have
Corollary 2.5.
For and , we have
The following lemma will be used in the pointwise estimate of the Poisson kernel. For , we denote
(2.7) |
For , it is defined in [16] for or , whose derivative estimates are given in [16, (62, 63)] for and [8, (2.11), (2.12)] for general . Now we give similar estimates for our generalized function .
Lemma 2.6.
For and integers with ,
(2.8) |
This lemma follows from , the estimate of in [8, (2.11)], and the inequalities
(2.9) |
The next two lemmas will be useful to generate the blow-up in the proof of Theorem 1.2.
Lemma 2.7.
For , we have
(2.10) |
where . The first term dominates the second when .
Proof.
The proof is similar to [3, Lemma 3.2]. We divide to three disjoint sets , and , which are defined by
and . We then split the integral into three terms as follows:
Since , we have
Thus, using , we have
Since , , we have
We decompose in the following way,
It’s not hard to get . Similar to , we have
Since , we have
Combining the above estimates, we arrive at (2.10). ∎
The following lemma is a generalized version of [3, (4.14)] and plays a crucial role in the proof of Theorem 1.2. Our lemma establishes both directions of the inequalities, while [3, (4.14)] only establishes the first inequality of (2.11), though only the first inequality is used later. Also our proof is more elementary and direct, without usage of the trace theorem of anisotropic Sobolev spaces. Recall that for function in homogeneous Besov space (see [1, Definition 2.15]), its norm is defined by
where is the Littlewood-Paley operator.
Lemma 2.8.
Let , , , where is the 1D heat kernel. We have that for ,
(2.11) |
Proof.
Notice that
(2.12) |
For fixed , we have
(2.13) |
Here is a constant. The first inequality is due to Littlewood-Paley decomposition and Minkowski’s inequality. The proof of the second inequality is similar to [1, Lemma 2.4]. Specifically, by definition and (2.12),
where is supported in some annulus with . Let be supported in some annulus with on the support of . By Young’s inequality and scaling technique, we can write the above equation as
Hence, what remains to show is
(2.14) |
We have
Note that no matter or . By using the fact
we arrive at (2.14) and hence (2.13). Therefore,
Note that the second inequality is by Hölder’s inequality , where
and since , see [1, Lemma 2.35]. Hence, the second inequality in (2.11) is valid.
To prove the first inequality in (2.11), we use the following identity from (2.12),
(2.15) |
Therefore, by Minkowski and Young’s inequalities and (2.14) (and the inequality above it), we have
Thus,
For the second inequality we have used Hölder inequality
where
For the third inequality we used again. Hence, the first inequality in (2.11) is proved. ∎
3. Derivation of Poisson kernel
In this section we derive the formula of Poisson kernel of system (1.1)–(1.2). We assume , and ( if ) are smooth functions satisfying (1.1)–(1.2) and vanishing sufficiently fast near infinity. We will use Fourier transform to calculate the Poisson kernel, which allows us to express and in terms of .
When , it is reasonable to assume , vanish sufficiently fast near infinity, which makes the Fourier transforms of and to be functions instead of distributions. However, we cannot expect pressure to vanish at infinity when , since and the fundamental solution for has no spatial decay (see also the first term in the RHS of (3.13)). Hence, the case is a bit subtle, and we need to treat it separately. Notice that , satisfy the Stokes equations (1.1) and Navier BC (1.2) with boundary value . We can then express and in terms of for . Hence, we can get the expression of and in terms of by the fact that and vanish at infinity. The formula of pressure is subject to change up to a function . In the end we will get the same form of the Poisson kernel for both and .
We show the derivation of the Poisson kernel here for and is analogous using the above observation. We imitate Solonnikov’s treatment of the velocity for Stokes system in [17], and decompose and into two parts
(3.1) |
Then is solenoidal and solves the heat equation
while is harmonic
Taking Fourier transform on the above heat equation and Laplace equation, we have
where are defined in (2.2). Solving the above ODEs in and removing the exponentially growing part, we obtain
for some function and . (Note and are different and are notations of [17].) By (3.1),
(3.2) | ||||
(3.3) |
Next, taking Fourier transform of and the Navier boundary condition (1.2), we get that
and for ,
We solve the above system and get
(3.4) | ||||
Substituting the above expression back to (3.2), we get that for ,
(3.5) |
Taking inverse Fourier transform of equations (3.3) and (3.5), we get and for . Observe that can be split into two parts, the part with and the remaining one
(3.6) |
where
(3.7) |
Hence, it seems natural to split the velocity and pressure into two parts, where the first part is the same expression with , and the second part is the difference. Namely,
(3.8) |
where for ,
(3.9) |
and
(3.10) |
In the following Lemma 3.1, we give the expression of the Poisson kernel of case in physical variables by taking inverse Fourier transform of equations (3.9) and (3.10). Derivation of , will be given before Lemma 3.3.
Lemma 3.1.
Proof.
The main tools are Lemma 2.1 to Lemma 2.5. From (3.9) and (3.7), for , we have
Note that in the second step above we use the following equality
and then apply . This trick will be used repeatedly in the later calculation.
For and , we have
(3.14) |
Now we derive the expressions of and defined in (3.8). Denote for simplicity. Using the facts
and the splitting of in (3.6), we have
(3.15) |
and for ,
(3.16) |
and
(3.17) |
The following Lemma will be useful in later calculation.
Lemma 3.2.
(3.18) | |||
and
(3.19) |
Proof.
Lemma 3.3.
Proof.
We next obtain the expression of by inverting , the factor of in (3) and (3). For and , by (3.19) when , we have
Using Lemmas 2.1, 2.4 and equations (3.18) and (3.28), we obtain (3.24). For and , by (3.19) when , we have
Using Lemma 2.1–Corollary 2.5 and equations (3.18) and (3.28), we obtain (3.26). ∎
Lemma 3.4.
Proof.
For , we have that and . By (3.5) we have for ,
where is defined in (3.4). By dominated convergence theorem, we have that in . Denote
It’s not hard to prove , and hence is bounded. Notice that, using ,
We obtain that if ,
(3.30) |
By dominated convergence theorem, we have that converge to in . Hence, we have in . Similarly, we can prove in . ∎
4. Estimate of Poisson kernel
In this section, we give upper bound estimates of the Poisson kernel derived in Lemma 3.1 and Lemma 3.3, and extend its usage to non-smooth boundary value .
Lemma 4.1.
Let . For and ,
(4.1) |
For and ,
(4.2) |
For and ,
(4.3) |
Remark 4.2.
When we evaluate the velocity given by (3.8) and Lemmas 3.1 and 3.3, with are more likely the source of singularity for high-order derivatives, because one factor of becomes in (4.3). Moreover, comparing (4.3) with the estimate of the Golovkin tensor in [16] or [9, (2.20)],
the only difference between the estimates of and is the exponent of .
Proof.
Next, for , , given in (3.24)-(3.25), we estimate –. We have
Since
by Lemma 2.6 and
(4.6) |
we have that for
By integration by parts in variable (to gain decay in when we apply Lemma 2.6), we have
Therefore, by Lemma 2.6 and (4.6), we have that for ,
Note that in the first line above we have used . For the two double integrals, we have first integrated in using (4.6), and then integrated in using .
Now for given in (3.26)-(3.27), we estimate –. By Lemma 2.6 and (4.6), we have that for
(4.7) |
Analogous to the estimates of , by integration by parts in variable , we have
(4.8) | ||||
By Lemma 2.6 and (4.6), we have that for ,
(4.9) |
By integration by parts in variable , we have
By Lemma 2.6 and (4.6), we have that for ,
(4.10) |
Therefore, by (3.26), we have that for ,
(4.11) |
and for ,
(4.12) |
These show (4.3). ∎
Lemma 4.3.
Let and . For all , we have
(4.13) |
where is independent of . Hence the convolution has a unique extension to all that satisfies (4.13).
Proof.
Recall is given by (3.12). It suffices to show (4.13) for . It is easy to check that
(4.14) |
where the constant is independent of . As for , it’s not hard to see . Thus,
Hence, by Young’s inequality, we obtain (4.13) for .
For , by [18, page 29, Theorem 1], we can prove
(4.15) |
As for with , by (3.14), we have
which is bounded uniformly in . Thus, the inequality
(4.16) |
holds for . By the estimate (4.5) we can verify the conditions for [19, page 19, Theorem 3] are fulfilled, which shows that (4.16) holds for . Using duality argument we can get that (4.16) holds for . Combining (4.16) and (4.15), we arrive at (4.13). ∎
Lemma 4.4.
Suppose , , and is constructed through the Poisson kernel by (1.6). Denote . For any compact set , we have
(4.17) |
Moreover, is spatially differentiable to any order when .
Proof.
Proof of Theorem 1.1.
First suppose . Similar to (4.18), , are smooth functions for both variables and . Similar to the proof of Lemma 5.2, , ( if ) vanish sufficient fast near infinity. Since are given in (3.5) and (3.3), it is not hard to verify they satisfy Stokes equations in Fourier space. Hence , satisfy Stokes equations (1.1) in physical space. By Lemma 3.4 we know satisfies Navier boundary condition (1.2). Hence, is a weak solution satisfying (1.7). Then using Lemma 4.4 and a density argument, we get the general case for . ∎
5. Blow-up of the second derivative in
In this section, we will prove Theorem 1.2. In the following we assume , since blow-up in implies blow up in . Using a similar idea as that in [3], we set the boundary value
for scalar function , where is a cutoff function
(5.1) |
and
(5.2) |
There is an example of function satisfying (5.2) in [3, Appendix C]. We will take for , and with for .
By Theorem 1.1, a velocity field satisfying Stokes equation (1.1) and Navier BC (1.2) with the given boundary data is
(5.3) |
for , where and are defined in (3.12) and (3.26). By (3.26), we set
(5.4) |
where for ,
(5.5) |
We give some identities which will be often used in this section. By the fact
(5.6) |
and (4.8) for , we obtain that for defined in (3.27),
(5.7) |
The first identity in (5.7) will not be used.
Lemma 5.1.
For defined in (5.3), we have , when and .
Proof.
We first consider the case .
First, we estimate , . Notice that by (5.6),
(5.8) |
and by (4.5),
(5.9) |
By (4.14), (5.8)–(5.9) and integration by parts of the horizontal derivatives , , it is easy to verify
(5.10) |
Next we estimate , . Using (5.7) and integration by parts of the horizontal derivatives , , we get
(5.11) |
By estimates (4.7), (4.9) and (4.10), the RHS of the above inequality is bounded, i.e.,
(5.12) |
Analogously for , we have
(5.13) |
where is defined in (5.5). Now we look at normal derivative of , ,
We first integrate by parts in variable to get
(5.14) |
where
We can obtain that by (4.7),
(5.15) |
by Lemma 2.6,
(5.16) |
and by Young’s convolution inequality, (4.14) and (4.6)
(5.17) |
Therefore, we have
(5.18) |
Hence, we get that for ,
(5.19) |
(5.20) |
Combining (5.10), (5.12), and (5.20), we have proved Lemma 5.1 for .
Lemma 5.2.
For defined in (5.3), we have that for and ,
(5.22) |
Proof.
Thus, by Lemma 5.1 and 5.2, we get the finite global energy (1.9) of for . For , by moving the derivative in to in (5.3), we are able to get a stronger decay estimate, for and ,
(5.27) |
Thus, it has finite energy (1.9). We leave the details to the interested readers. Now, we show the blow-up of second derivative of the solution in , .
Lemma 5.3.
For defined in (5.3) and , we have
(5.28) |
Proof.
For , we can obtain that by (4.1),
(5.29) |
and by estimate (4.9) and (4.10), both with ,
(5.30) |
Next we work on
We still use the decomposition (5.14) to get
For , by (5.19), we know that for ,
(5.31) |
Moreover, for , using the fact that as , we have
Thus, by integration by parts of variable , we have
where
By Lemma 2.6 and (4.6), we have
(5.32) |
Therefore, by (4.4), (5.25), (5.32), and for , we are able to obtain
(5.33) |
Next, we denote the leading blow-up term from :
(5.34) |
where
using . By Lemma 2.7, we can decompose into
where
is the 1D heat kernel, and
For ,
Hence, we have that for ,
(5.35) |
Next, we factor as
(5.36) |
where
(5.37) |
for ,
and for ,
Actually for , we have that . Hence for we have , and for we have . Hence we have .
Since is supported in , is defined for , and vanishes for . We have for ,
and for ,
This implies that for , using (4.6),
(5.38) |
Note that both inequalities above are not true when . By Lemma 2.8, we have
(5.39) |
With the aid of (5.36), (5.38), (5.39), and the fact that when , we conclude that
(5.40) |
Hence
By the above unboundedness and (5.29)-(5.31), (5.33), (5.35), we obtain that
(5.41) |
6. Shear flow example
In this section, we prove Theorem 1.4. We follow Seregin-Šverák [13] and look for shear flow solutions of the Stokes system (1.1)–(1.2) in with Navier boundary value in the form
(6.1) |
where . It’s easy to see that the convection term is zero, so it is also a solution of Navier–Stokes equations. Denote , and we can reformulate the original equations as
(6.2) |
For the heat equation with Robin BC [10], the solution is given by
(6.3) |
using its Green function
Here is 1D heat kernel. Direct calculation gives
(6.4) | ||||
Since
the boundary condition is satisfied. It is easy to check in (6.3) is a smooth solution (up to the boundary ) of heat equation (6.2).
By (6.4), we get
and
Using integration by parts of the derivative , we can obtain an equivalent formula of (6.3),
(6.5) |
Notice that
so by Young’s convolution inequality,
for . Since is dense in for , by the density argument it is easy to check that for any with , in (6) satisfies the weak form of (6.2), and hence given by (6.1) satisfies the weak form of Stokes systems (1.7) with Navier boundary value .
Lemma 6.1.
Proof.
The above lemma proves Theorem 1.4 for the , and the blow-up in follows from blow-up in .
Remark 6.2.
Acknowledgments
We warmly thank K. Kang for fruitful discussions on his paper [3] with T. Chang. We also thank Rulin Kuan for reference [12] for Remark 2.2. Hui Chen was supported in part by National Natural Science Foundation of China under grant [12101556] and Zhejiang Provincial Natural Science Foundation of China under grant [LY24A010015]. The research of both Liang and Tsai was partially supported by Natural Sciences and Engineering Research Council of Canada (NSERC) under grant RGPIN-2023-04534.
References
- [1] H. Bahouri, J.-Y. Chemin, and R. Danchin. Fourier analysis and nonlinear partial differential equations. Springer, Heidelberg, 2011.
- [2] T. Chang and K. Kang. On Caccioppoli’s inequalities of Stokes equations and Navier–Stokes equations near boundary. J. Differential Equations, 269(9):6732–6757, 2020.
- [3] T. Chang and K. Kang. Local regularity near boundary for the Stokes and Navier–Stokes equations. SIAM J. Math. Anal., 55(5):5051–5085, 2023.
- [4] H. Chen, S. Liang, and T.-P. Tsai. Gradient estimates for the non-stationary Stokes system with the Navier boundary condition. Commun. Pure Appl. Anal., 2023.
- [5] H. Dong, D. Kim, and T. Phan. Boundary Lebesgue mixed-norm estimates for non-stationary Stokes systems with VMO coefficients. Comm. Partial Differential Equations, 47(8):1700–1731, 2022.
- [6] S. Gustafson, K. Kang, and T.-P. Tsai. Regularity criteria for suitable weak solutions of the Navier–Stokes equations near the boundary. J. Differential Equations, 226(2):594–618, 2006.
- [7] K. Kang. Unbounded normal derivative for the Stokes system near boundary. Math. Ann., 331(1):87–109, 2004.
- [8] K. Kang, B. Lai, C.-C. Lai, and T.-P. Tsai. Finite energy Navier–Stokes flows with unbounded gradients induced by localized flux in the half-space. Trans. Amer. Math. Soc., 375:6701–6746, 2022.
- [9] K. Kang, B. Lai, C.-C. Lai, and T.-P. Tsai. The Green tensor of the nonstationary Stokes system in the half space. Comm. Math. Phys., 399(2):1291–1372, 2023.
- [10] J. B. Keller. Oblique derivative boundary conditions and the image method. SIAM J. Appl. Math., 41(2):294–300, 1981.
- [11] F. Lin, Y. Sire, J. Wei, and Y. Zhou. Nematic liquid crystal flow with partially free boundary. Arch. Ration. Mech. Anal., 247(2), 2023.
- [12] L. Schwartz. Théorie des distributions. Hermann, 1966.
- [13] G. Seregin and V. Šverák. On a bounded shear flow in a half-space. Zap. Nauchn. Sem. S.-Peterburg. Otdel. Mat. Inst. Steklov. (POMI), 385:200–205, 2010. J. Math. Sci. (N.Y.) 178 (2011), no. 3, 353–356.
- [14] G. A. Seregin. Some estimates near the boundary for solutions to the nonstationary linearized Navier–Stokes equations. Zap. Nauchn. Sem. S.-Peterburg. Otdel. Mat. Inst. Steklov. (POMI), 271:204–223, 2000. J. Math. Sci. (N.Y.) 115 (2003), no. 6, 2820–2831.
- [15] G. A. Seregin. A note on local boundary regularity for the Stokes system. Zap. Nauchn. Sem. S.-Peterburg. Otdel. Mat. Inst. Steklov. (POMI), 370:151–159, 2009. J. Math. Sci. (N.Y.) 166 (2010), no. 1, 86–90.
- [16] V. A. Solonnikov. Estimates for solutions of a non-stationary linearized system of Navier-Stokes equations. Trudy Mat. Inst. Steklov., pages 213–317, 1964.
- [17] V. A. Solonnikov. Estimates for solutions of nonstationary Navier–Stokes equations. Journal of Soviet Mathematics, 8(4):467–529, 1977.
- [18] E. M. Stein. Singular integrals and differentiability properties of functions. Princeton Mathematical Series, No. 30. Princeton University Press, Princeton, N.J., 1970.
- [19] E. M. Stein. Harmonic analysis: real-variable methods, orthogonality, and oscillatory integrals, volume 43 of Princeton Mathematical Series. Princeton University Press, Princeton, NJ, 1993. With the assistance of Timothy S. Murphy, Monographs in Harmonic Analysis, III.