Curvature II: Spacetime

By Scott Church – Guest Blogger

In the first installment of this series, we explored the nature of curved spaces and introduced ourselves to some of the mathematical tools needed to describe how length, breadth, and height can be curved without higher dimensions to “curve into.” In the interest of keeping our exploration as intuitive as possible, we began with the Euclidean geometry we learned in high school and explored curvature from the vantage point of time as we experience it—a universal history that is the same for all of us and independent of the spatial stage on which our lives unfold. Today we will explore the nature of time and its relationship to space and discover (spoiler alert!) that in fact, it is neither separate from space nor absolute—not only can length, breadth, and height be curved, duration can be as well. The universe we inhabit is one of curved spacetime.

Special Relativity

The Newtonian physics we learned in high school presumes absolute three-dimensional space and time. In the low gravity and velocity world we live in, that is how we experience them. But intuitive as this may seem to us, there are hints that something is amiss. That physics also taught us that the speed of light c is a universal constant that can be derived from Maxwell’s equations. And as we saw in Part I, the laws of physics, including c, must be invariant for all observers stationary or moving. Pause for a moment and reflect on what this implies. If I am standing beside a highway and you drive by at 50 mph, that is the speed I will observe. In the car, you will see yourself as stationary and the world passing you at 50 mph in the opposite direction, including me. Another driver doing 70 mph in the fast lane will pass me at that speed and you at 20 mph. But Maxwell’s equations will remain true and invariant for all observers, so if a beam of light is shined in the same direction, it will pass all three of us at the same speed. How is this possible?

Imagine that you are now the one who is stationary, and I fly past you in a fighter jet at a speed v of 3600 mph, (or one mile/sec for round numbers) carrying a clock that is in sync with an identical clock of yours. As I pass you, it emits a pulse of light in your direction at time t_1 which reaches your eye after travelling a distance d_{t1} (Figure 1). One second later at t_2, a second pulse is emitted, but I will have flown one mile further so that pulse must travel a distance d_{t2} before it reaches your eye. My clock will be ticking at the same rate in my reference frame as yours is for you, but the seconds you observe on my clock will be longer because the second pulse you receive from it must travel further at the same speed c to reach your eye than the first one did. Your experience will be that my time runs slower for you than it does for me. And for the same reason, your clock will be running slower for me than it is for you.

 

Figure 1

As for distances, the length of my jet will be measured by the time it takes a pulse of light to travel from the nose (A) to the tail (B) at speed c (Figure 2). In my reference frame that will be given by,

L = c\Delta t_{ba}                    [Eqn. 1]

 where t_{ba} is my proper time (that is, the time measured by a clock at rest in my reference frame).

Figure 2

In your reference frame, the pulse of light will take a time \Delta t^{'}_{ba} to travel the length of my jet. However, while the pulse is in transit, point B will have moved forward a distance v \Delta t{'}_{ba} so the pulse will arrive at point B_2 instead (Figure 3),

Figure 3

And you will observe the length of my jet to be the distance between A and B_2, or,

L^{'} = (c - v)\Delta t^{'}_{ba}                    [Eqn. 2]

Not only will you see the pulse travelling a shorter distance that me, the time \Delta t^{'}_{ba} will also be less than the \Delta t_{ba} I observe because time is running slower for you than for me. The length L^{'} you observe for my jet will be smaller than the length L I observe, and you will see me and my jet as though we were compressed in the direction of travel.

Thus, we arrive at one of the foundational principles of special relativity; Space and time are neither absolute nor independent of each other. They’re united in a single spacetime manifold whose metric contains an underlying symmetry that preserves Maxwell’s equations and c for all observers. And this manifold is not simply a map of locations and distances—it’s a frame-independent history of events for every location within it.

In Part I we saw that in the flat Newtonian universe of our experience, time is absolute and independent of space. All observers experience it the same, and spatial geometry is Euclidean with the interval between any two points is given by the Pythagorean Theorem,

 ds^{2} = dx^{2} + dy^{2} + dz^{2}                    [Eqn. 3]

In spacetime, however, this is no longer the case. Now we have a collection not of points, but events that reflect the histories of each spatial point within it. The interval no longer defines the distance from here to there; It defines here and now, to there and then. Accounting for this in our metric tensor won’t be as simple as it may sound. As we’ve seen, the speed of light must remain the same for all observers whether stationary or moving in any reference frame. And the relative motion slows time down and compresses space until both reach zero at the speed of light. From our vantage point, a photon’s reference frame is a single event with a zero-length interval, so our interval must include time with a sign opposite to that of space. After multiplying time by c to convert it to equivalent distance units, this gives,

ds^{2} = dx^{2} + dy^{2} + dz^{2} - (ct)^{2}                    [Eqn. 4]

Which adopting the usual (though not strictly necessary) convention of making time the first, or zero component, results in the spacetime metric tensor,

                              [Eqn. 5]

The diagonal terms expressed as a tuple, [-1, 1, 1, 1], is known as the metric’s signature. In differential geometry (the branch of mathematics that generalizes the geometry we learned in high school to all types of curves spaces), a continuous N-dimensional manifold that has a well-defined and positive-definite metric tensor at all points (not all mathematically possible ones do) is referred to as Riemannian. That is, a flat 4-D Riemannian metric is one that for every point on it, infinitesimal displacements in the locally flat tangent plane have a metric signature of [1, 1, 1, 1]. A universe with Euclidean geometry and absolute time would have this metric everywhere. But in a universe constrained by special-relativity the interval can be zero as well as positive, so the metric is non-degenerate rather than positive-definite. Manifolds of this type are referred to as pseudo-Riemannian.

In Part I, we conducted a geometric thought experiment in which we traversed a closed triangular path through the flat space of an observer named Freddy, and another through the curved space of an observer named Cathy along geodesics (paths that reflect the shortest distance between any two points). In each, we carried one vector with us while leaving an identical parallel copy of it behind and upon returning to point A. When we did this in Freddy’s flat space we found, not surprisingly, that after completing the journey the two vectors were still parallel to each other. But after the same journey through Cathy’s curved space, we discovered that the vector we carried with us was no longer parallel to the one we left behind even though both were still pointing in the same direction (globally south), and we had travelled a shorter distance that still encompassed a larger area. We introduced some mathematical concepts that allowed us to define a covariant derivative 1 to describe the rate of change of the vector we carried with us along our path s^{\sigma},

\nabla_{\mu}s^{\sigma} = \partial_{\mu}s^{\sigma} + \Gamma^{\sigma}_{\mu\nu}s^{\nu}                    [Eqn. 6]

The first term on the right is the usual vector calculus gradient along the direction of travel. The second term, however, introduced a new object, the Christoffel symbol, that allowed us to map changes in the underlying tangent plane containing s^{\sigma}, itself onto local coordinate systems within it as we traversed the path. Integrating this derivative along our path would then fully capture the changes in our mobile vector with respect to its stationary twin we left behind.

That exercise, however, traversed a path through Cathy’s curved two-dimensional space, so equation 6 described distances and directions only. Had we included time in her curved universe, the path we walked would have been a trajectory of motion with history, and upon arriving back at A we would have found that our mobile vector was now older or younger than its stationary copy as well. In curved space, geodesics are the shortest distance between points—here and there. But in spacetime they are histories that reflect the shortest path, stationary or moving, between here and now, and there and then. As such, they define an equation of motion for the trajectory an object will follow when no forces are acting on it.

In a flat spacetime like Freddy’s, an object left to itself will remain stationary or move at constant velocity, so its geodesic will be a straight line whose slope will be the constant speed it is moving at. If one or more forces act on the object it will accelerate, and its history will follow a curved path whose velocity changes from moment to moment. We can derive the equation of motion for this by using equation 5 to derive the second order time derivative along ds^{\sigma}, to equate the acceleration produced by a force it to its strength divided by the object’s mass. In his flat spacetime, a single unvarying tangent plane spans the entire universe, so the Christoffel term will vanish, leaving us with,

\frac{\partial^2 s^{\sigma}}{\partial t^2} = \frac{F}{m} = 0                    [Eqn. 7]

Which we will recognize as a geodesic equation of motion for Newton’s second law that we learned in high school.

In Cathy’s universe things are different. There, geodesics are curved so the Christoffel term will generally be non-zero, and her equation of motion will be given by,

\frac{\partial^2 s^{\sigma}}{\partial t^2} + \Gamma^{\sigma}_{\mu\nu}\frac{\partial s^{\mu}}{\partial t^2}\frac{\partial s^{\nu}}{\partial t^2} = 0                    [Eqn. 8]

Notice that in curved spacetimes like hers, the second term on the left will be non-zero even in the absence of forces, so the first term will be as well. Left to themselves, objects in a curved spacetime will experience freefall along accelerating trajectories.

Which brings us to the next topic…

General Relativity

The other hallmark of our high-school physics lessons was Newtonian gravity. In a universe of flat space and absolute time like Freddy’s, gravity is an attractive force between objects whose strength is a function of their masses and the distance separating them. Specifically, the gravitational force F_g between two objects with masses m_1 and m_2 is given by,

F_g = g_c\frac{m_1m_2}{r^2}                    [Eqn. 9]

Where r is the distance between their centers of mass and g_c is the universal gravitational constant we also learned in our high-school physics classes.

For centuries this understanding of gravity has served us well in regions of low mass, velocity, and distance, and still does. I spent twenty years as an aerospace engineer designing commercial jet aircraft structures, and the aircraft my colleagues and I applied these principles to still have exemplary safety and performance records. But even so, physicists have long been troubled by the idea of “spooky action at a distance” forces. How can objects interact with each other invisibly over large distances? On the other hand, we can put it differently by saying that gravity causes objects with mass to accelerate toward each other at a rate given by their masses and the distance separating them, and as we saw above, freefall acceleration is a consequence of spacetime curvature. Jumping the gun, we also know that mass and energy are equivalent (hence Einstein’s celebrated E = mc^2) and moving objects with mass have a kinetic energy that is a function of their momentum and mass (K = p^2/2m). This raises an interesting question…

What if gravity isn’t a force at all, but simply a local manifestation of spacetime curvature due to mass, energy, and momentum?

If this is true, then we would expect that two objects of differing mass in the field of a third object of much larger mass (like the earth, for instance) would experience the same freefall acceleration toward it—essentially, that the “force” F_g the gravitational field exerts on their differing small masses would result in the same acceleration for both,

\frac{F_{g1}}{m_1} = \frac{F_{g2}}{m_2}                    [Eqn. 10]

And this would be the same acceleration that would result from an equal but non-gravitational force (e.g. - the thrust produced by a rocket engine). As you’ve probably guessed by now, this is the case. Gravitational mass and inertial mass are indistinguishable from each other, and freefall accelerations induced by the former are a consequence not of any “spooky action at a distance” force, but of the local spacetime curvature created by its presence. This identity, known as the equivalence principle, is the heart and soul of general relativity. Throw a pebble into a pond and watch it arc through the summer sky before splashing down, and you are literally seeing the curvature of length, height, breadth, and duration where you’re standing because of the mass of the earth beneath your feet! 2

And once again, if spacetime curvature is caused by mass, energy, and momentum, we can ask ourselves how this could be captured mathematically. As in Part I, a formal derivation of the relationship between the two is beyond the scope of an introduction to the topic, but we can introduce the types of mathematical objects needed and how they relate to each other. The first thing we need is an object that describes curvature. Like the terms introduced so far, it will need to capture the change in angles over infinitesimal displacements from any reference frame we view it from, so it will need to be a covariant or contravariant tensor. And since we want it to describe curvature specifically rather than displacements, it will be a function of the Christoffel symbols that describe how they change when we walk a parallel transport path (or more properly, a function of their first derivatives, or rates of change). To unambiguously capture this, we will have to carry a four-vector ds (that is, a vector in three spatial dimensions plus time) around an enclosed path for which all the interior angles are orthogonal to each other (locally 90 degrees). Previously, we were able to do this with a triangular path in Cathy’s space because for clarity of the underlying principles we presumed it to be spherically curved, but that won’t be true of curved spacetime in general. So, now we must carry our four-vector along a four-legged parallel transport path (presumed to be infinitesimally small for a local curvature description), again preserving its local orientation at every point, as shown in Figure 4 (Wikimedia, 2015).

Figure 4

Upon returning to our starting point, we will have a function that describes how each of the four components of ds changed with respect to the others for each of the four legs of the journey. As such it will be a tensor with four indices (rank 4) each of which covers four dimensions, so it will have 4^4, or 256 components. This tensor, known as the Riemannian curvature tensor R^{\mu}_{\nu\rho\sigma}, fully describes the actual curvature of spacetime at every point on the manifold. It can be specified in covariant or contravariant terms, but since it captures how a contravariant vector is affected by local covariant curvature, it’s customary to express it with one “upstairs” index and three “downstairs” ones, as shown here.

Before going any further, there are two related tensors we’re going to need (why will become apparent shortly). In Part I we discussed how a tensor object defined by N indices can be “contracted” to fewer indices by projecting one or more of the index’s components onto the others—in essence, “averaging” it into the remaining ones. For a tensor expressed in covariant form for all indices, we do this by multiplying it by the contravariant metric tensor in one or more of its indices. Contracting the Riemann tensor in this manner for two of its four indices gives,

 g^{\rho\sigma}R_{\mu\nu\rho\sigma} = R_{\mu\nu}                    [Eqn. 11]

The resulting tensor, R_{\mu\nu}, is known as the Ricci tensor. Contracting it again on both of its indices yields the Ricci scalar, R. These have different physical interpretations. The Ricci tensor describes the rate of change of an infinitesimal element of spacetime volume along ds due to tidal forces. That is, as we move through spacetime along a group of infinitesimally separated parallel geodesics, it describes how an element of volume between them changes in each direction. The Ricci scalar, on the other hand, gives a non-dimensional measure of how the overall enclosed volume itself changes.

Next, we need a tensor object that describes the mass, energy, and momentum we suspect to be curvature’s source. That tensor (which we won’t make any attempt to formally derive here), is known as the stress energy momentum tensor, T^{\mu\nu}. Its components are defined in a manner similar to those of the metric tensor, g_{\mu\nu}, but using momentum density four-vectors (momentum density in three spatial dimensions plus energy density, which can be thought of as “momentum” in time for a stationary object). Because its momentum density components are vectors, it is customary to express it in contravariant form (indices “upstairs”). The first index (\mu) gives the four-momentum components being considered, and the second (\nu) gives the direction it is being compared to. The physical significance of its components is as shown in Figure 2 (Wikimedia, 2013).

 

Figure 5 – The Stress Energy Momentum Tensor

With these tools in hand, we can proceed with our investigation of how mass, energy, and momentum curve space and time, but there are still a few constraints we need to account for.

First, the stress energy momentum tensor is rank 2 but the Riemann curvature tensor is rank 4 (that is, the former has two indices with 16 components, whereas the latter has 4 indices and 256 components), so we can’t just equate them to each other. Whatever effect T^{\mu\nu} has on curvature will have to manifest itself as a rank 2 curvature object as well—that is, it will have to be a contraction of the Riemann tensor that reflects the behavior we observe in gravity, so we want to know what sort of contraction will give us that.

We saw earlier that in the absence of forces, spacetime curvature manifests as acceleration. Strictly speaking, this applies only to point masses in the gravitational field of a much larger mass. For objects that have size and shape, the story changes. In Newtonian physics, the gravitational force between two masses varies inversely as the square of the distance between them (equation 9). So, if you are falling toward the earth feet first, your feet are being pulled harder than your head because they are closer to the earth’s center of gravity. Inasmuch as this is the low mass/energy/momentum limit of GR, the same will be true in curved spacetime as well. Likewise, your freefall into the earth’s gravitational well will be along a geodesic, and the deeper you go, the closer adjacent geodesics to your sides will be. Figure 3 (Wikimedia, 2008) shows what a gravitational well created by a mass as the bottom of the “pocket” looks like.3 The longitudinal lines are freefall geodesics with their steepness at each node being the strength of gravity there, and the squares enclosed by the grid can be thought of as shapes.

 

Figure 6 – Gravitational Well

Notice how falling into the well squeezes the latitudinal rectangles into increasingly longitudinal ones. In the earth’s relatively weak gravitational field compared to your size, the effect is too small to notice. But as you fall toward it, feet-first, you are being stretched and squeezed. This stretching and squeezing of large objects are tidal forces, and in the limit of a point mass, they reduce to simple freefall acceleration. Since in the most general terms, tidal forces are how curvature manifests, we would expect the stress energy momentum tensor to equate to a rank 2 tensor that describes them. And as we’ve seen, we have one… the Ricci tensor!

But we’re not out of the woods yet. There is one more constraint we need to honor; Another of the fundamental ones we learned in our high school physics, conservation of energy and momentum. Although neither is well-defined nor self-evidently conserved for the whole universe (or large regions of it), for locally flat inertial reference frames in the tangent planes of every point in it, both need to be conserved. This means that for every point on the manifold the divergence of the stress energy momentum tensor must be zero. That is,

\nabla_{\mu}T^{\mu\nu} = 0                   [Eqn. 12]

And here we have a problem… Tidal forces do not vanish in locally flat regions, and neither does the divergence of the Ricci tensor. If they did, falling through a black hole event horizon would be a lot less traumatic! So, our contracted curvature tensor object is going to need some tweaking.

Fortunately, the full Riemann curvature tensor itself gives us a way out. As it happens, its own internal consistency does require it to vanish locally; When curvature vanishes (as it must in local tangent planes) so does the curvature tensor. One consequence of this is that the sum of its divergences with respect to any three of its four indices must add to zero. That is,

\nabla_{\mu}R^{\mu }_{ \nu\rho\sigma} + \nabla_{\nu}R^{\nu}_{\mu\rho\sigma} + \nabla_{\rho }R^{\rho}_{\mu\nu \sigma} = 0                   [Eqn. 13]

This relationship is known as the second Bianchi identity (of which there are several). Again, we needn’t worry about its formal derivation here. But for our purposes, what matters is that with some mathematical gymnastics we can derive from it the contracted Bianchi identity,

\nabla_{\mu}R^{\mu\nu} = \frac{1}{2}\nabla_{\mu}g^{\mu\nu}R                   [Eqn. 14]

 Gathering terms gives,

\nabla_{\mu}\left ( R^{\mu\nu} - \frac{1}{2}g^{\mu\nu}R \right ) = 0                   [Eqn. 15]

And finally, by combining the Ricci tensor for tidal forces and the Ricci scalar for volumetric curvature, we have a tensor object we can equate to the stress energy momentum tensor that captures the spacetime curvature it induces while sharing with it a zero divergence that locally preserves conservation of energy and momentum. It’s customary to refer to the term in brackets as the Einstein tensor G^{\mu\nu}, from which we have,

G^{\mu\nu} = R^{\mu\nu} - \frac{1}{2}g^{\mu\nu}R = \kappa T^{\mu\nu}                   [Eqn. 16]

Where \kappa is a proportionality constant which again, we won’t derive here, but turns out to be,

\kappa = \frac{8\pi g_c}{c^4}                   [Eqn. 17]

And there you have it, Ladies and Gentlemen… an equation that relates mass, energy, and momentum to spacetime curvature, and therefore gravitation!

One final question remains. Technically, equation 16 is arbitrary to within an additive constant as well. When Einstein first derived this relationship, he realized that it predicted a universe that was necessarily expanding or contracting, and thus impermanent. The idea of a universe that wasn’t eternal was philosophically abhorrent to him, so he included a constant term on the left (typically denoted with the Greek letter \Lambda), multiplied by the metric tensor for consistency and sized to offset the expansion, thereby preserving a curved, but static and eternal universe. Later, when it was independently confirmed that the universe is in fact, expanding (a fascinating story in its own right!), Einstein retracted the constant calling it “the greatest mistake of my life.” But as it turns out, it wasn’t. It has since been discovered that the cosmological constant is not only real, but positive and causing the expansion of the universe to accelerate! The discovery was so striking that the leaders of the team who discovered it, Saul Perlmutter, Brian Paul Schmidt, and Adam Guy Riess were jointly awarded the 2011 Nobel Prize in physics.

So… combining equations 16 and 17 with all terms expressed as covariant (which is customary), and restoring the cosmological constant to its rightful place we have,

G_{\mu\nu} + \Lambda g_{\mu\nu} = \frac{8\pi g_c}{c^4} T_{\mu\nu}                   [Eqn. 18]

These are the celebrated Einstein Field equations that are the hallmark of general relativity. The terms on the left fully describe the geometry of spacetime for all observers at every point in the universe, and the term on the right describes the mass, energy, and momentum that produces that geometry.

This was meant to be an introduction to spacetime curvature, so we’ve arrived at them with some big leaps and little in the way of formality. Though at first blush they may seem daunting and difficult to wrap your mind around, the important thing for today is an understanding of what the terms in these equations mean, and why they must have the general forms they do to describe how length, height, breadth, and duration can be curved. For those who want to explore further, there any number of good introductions to general relativity for the layperson. One that I found particularly readable and informative was Clifford Will’s book Was Einstein Right – Putting General Relativity to the Test (1993), first published in 1986 when I was in grad school. If you feel ready to make the deep dive into the full formalism of general relativity, there are many textbooks on the subject. But if there is one that has stood for many years as the Bible of general relativity, it’s Misner, Thorne, and Wheeler's Gravitation (2017). It’s rigorous and will take some time to wade through, but it’s the best, and most thorough general relativity course I am personally aware of and has been since it was first published in 1973.

The psalmist tells us,

“The heavens are telling the glory of God; and the firmament[a] proclaims his handiwork. Day to day pours forth speech, and night to night declares knowledge. There is no speech, nor are there words; their voice is not heard; yet their voice goes out through all the earth, and their words to the end of the world.” – Psalm 19:1-4

When I gaze up at the nighttime sky, I see stars that are hundreds of light years away, many of which are surrounded by worlds, possibly even worlds not unlike my home. And I realize that I’m gazing upon those stars and worlds not as they are now in my reference frame, but as they were centuries ago. If I were to turn a large enough telescope on that sky I would see galaxies, quasars, nebulae, and a bewildering spectacle of other wonders, some of which are billions of years old and revealing themselves to me from a time long before humans or even our solar system existed. And if I filter their light through a spectrometer, I will see the fingerprints of their chemical constituents shifted increasingly toward the red the more distant they were, and I would realize that I was watching the universe grow—not as an expansion of matter into a pre-existing void, but literally the expansion of space and time themselves from a cataclysmic birth 13.73 billion years ago. I would see in that the glory of God and his handiwork…

And I would suspect, as J.B.S. Haldane did a century ago, the handiwork of God, where length, breadth, height, and duration are themselves clay in His artistic hands, is not only queerer than I suppose, but queerer than I can suppose.

Footnotes

1)  In Part I we introduced the nabla symbol on the left (\nabla_{\mu}), which in mathematics is known as the Laplace operator. It is a shorthand reference for the gradient (first derivative) in the direction of a vector defining the \mu coordinate system. That is, \nabla_{\mu} = \frac{\partial }{\partial x_0} + \frac{\partial }{\partial x_1} + \frac{\partial }{\partial x_2} + \frac{\partial }{\partial x_3} where the index \mu = 0, 1, 2, 3. This representation of a gradient in a particular direction is also referred to as the divergence.

2)  Interestingly, this isn’t just theoretical. Google and Apple map apps leverage first-order corrections for spacetime curvature near the earth’s surface to refine the accuracy of your location from raw GPS triangulated signals. General relativity is literally why your phone knows your location to within a couple hundred feet or so rather than one or two city blocks!

3)  Strictly speaking, this is a 2-D gravitational well with absolute time rather than a true 4-D gravitational which would include time. But for the current purpose, it suffices to illustrate the point.

References

Misner, C.W., Thorne, K.S. & J.A. Wheeler. 2017. Gravitation. Princeton University Press (Oct. 24, 2017). ISBN-10: 9780691177793, ISBN-13: ‎978-0691177793. Online at https://www.amazon.com/Gravitation-Charles-W-Misner/dp/0691177791/ref=sr_1_1?crid=1OKXLNQA5YVAR&keywords=gravitation&qid=1694219167&sprefix=gravitation%2Caps%2C253&sr=8-1&ufe=app_do%3Aamzn1.fos.18630bbb-fcbb-42f8-9767-857e17e03685.  Accessed Oct. 9, 2023.

Wikimedia. 2008. Image courtesy of AllenMcC. Based on the work of Bamse, and Melchoir, CC BY-SA 4.0, Mar. 2, 2013. Online at https://commons.wikimedia.org/wiki/File:GravityPotential.jpg. Accessed Oct. 9, 2023.

Wikimedia. 2013. Image courtesy of Maschen. Based on the work of Bamse, and Melchoir, CC BY-SA 4.0, Mar. 2, 2013. Online at https://commons.wikimedia.org/w/index.php?curid=24940142. Accessed Oct. 9, 2023.

Wikimedia. 2015. Image courtesy of IkamusumeFan, CC BY-SA 4.0, Jan. 1, 2015. Online at https://commons.wikimedia.org/w/index.php?curid=2615879. Accessed Oct. 9, 2023.

Will, C.N. 1993. Was Einstein Right? - Putting General Relativity to the Test. Basic Books; 2nd edition (June 2, 1993). ISBN-10: ‎0465090869; ISBN-13: ‎978-0465090860. Online at https://www.amazon.com/Was-Einstein-Right-Putting-Relativity/dp/0465090869/ref=sr_1_1?crid=TOG1ZAWGPF20&keywords=was+einstein+right&qid=1696883510&sprefix=was+einstein+right%2Caps%2C171&sr=8-1. Accessed Oct. 9, 2023.

Posted in Physics | 10 Comments

Curvature I: Space

My own suspicion is that the Universe is not only queerer than we suppose, but queerer than we can suppose. – J.B.S. Haldane (Possible Worlds and Other Papers, 1927)

I was born hopelessly curious and under the tutelage of a nurturing teacher and parents who surrounded me with books, I fell in love with physics in the 2nd grade—when all my friends were enthralled with Batman, jets, and G.I. Joe. What drew me to it was the wonder of mysteries I couldn’t wrap my budding mind around, and chief among these was the notion that space, and time could be curved. I remember pouring over my parent’s Time-Life encyclopedia set which among other things, contained a full-color plate titled “Three kinds of space” featuring gridded surfaces shaped like a sphere, a pancake, and a saddle labeled +1, 0, and -1 respectively (the Friedmann constants, although of course, I didn’t know that then). I remember gazing at them struggling to understand… How can length, breadth, height, and duration be bent…? What does that even mean…? The question became even more mind-numbing when I later discovered that there can be spaces with more than three or four dimensions—indeed, an infinite number of dimensions—and these can all be curved as well. It wasn’t until well into graduate school that I started to get a shaky footing in that recondite landscape.

As three-dimensional beings, most of us grasp curvature visually. We can see curved lines and sheets against the backdrop of three dimensions because they bend into the other dimension/s. But how can three-dimensional space (or more properly four-dimensional space-time) bend when there are no other dimensions to bend into? The key to understanding this is to approach the question not by trying to visualize higher-dimensional spaces, but by exploring them with a mathematically based thought experiment physicists and mathematicians refer to as parallel transport. Let’s introduce two explorers: Flat Freddy who lives in a two-dimensional flat universe, and his sister Curved Cathy who lives in a curved one. For them, there is no third dimension much less any higher ones.

Parallel Transport

Let’s start with Freddy, placing him at the vertex of a triangle with two parallel vectors oriented along his direction of travel, one red and one green (Figure 1).

Figure 1

Now, let him go for a walk around the triangle’s perimeter in the direction the vectors are pointing, leaving the green vector behind, and taking the red one with him while ensuring that for the entire journey it remains oriented in the same direction (as we will soon see, this matters). Completing the first leg of the journey, he arrives at point B (Figure 2) with his red vector still parallel to the green one, and unchanged from its original orientation (light red).

 

Figure 2

Then, let’s have him journey an equal distance to the right at a 90-degree angle. When he arrives at point C, his red vector is still parallel to the green one and its previous orientations (Figure 3).

 

Figure 3

Finally, let’s take Freddy back home and reunite his two vectors. When Freddy checks his compass, he sees that point A is to his left and back at a 45-degree angle to the BC leg he just covered. When he arrives home again, he finds that his red and green triangles are still parallel to each other, exactly as they were when he began, and remained throughout his trek (Figure 4).

 

Figure 4

Getting his map out, Freddy sees that his journey traversed a right triangle with two 45-degree angles, the final leg of which covered a distance given by the Pythagorean Theorem,

\overline{AC} = \sqrt{({\overline{AB})^{2}} + ({\overline{BC})^{2})            [Eqn. 1]

And enclosed an area given by,

A = \frac{X^{2}}{2}           [Eqn. 2]

Where X is the length of \overline{AB} (or \overline{BC}). No surprises here. This is exactly what earth-bound three-dimensional creatures like us would expect.

Parallel Transport in Curved Space

Now, let’s have Cathy take the same journey in her universe. For clarity’s sake, let’s assume her universe is spherical with a “radius” that will better illustrate the outcome (more on why that word is in quotes soon). Like Freddy, we’re going to have her walk a triangular path beginning at point A with parallel red and green vectors, both tangent to the straightest path from point A to point B (Figure 5). As before, she will leave the green vector behind while carrying the red one with her, keeping it oriented in the same direction throughout. This time however, things are going to be a little more subtle. In Freddy’s universe the meaning of “straight” is clear enough. But as we will soon see, in Cathy’s this term will require a more precise definition.

 

Figure 5

When she completes the first leg of her journey at point B, her red vector hasn’t changed orientation. It is still pointing straight ahead, tangent to her path of travel (Figure 6).

Figure 6

Following in Freddy’s footsteps, she then journeys an equal distance to the right at a 90-degree angle, arriving at point C with her red vector still unchanged in direction (Figure 7).

 

Figure 7

Cathy has now travelled the same route from point A to point C that Freddy did in his universe and covered the same distance getting there. But now, something is amiss. When she checks her compass, she finds that point A isn’t to the left of her BC leg and 45 degrees back. Home is now 90 degrees to her left. Even more strangely, upon arriving home (Figure 8) she sees that her red vector is no longer parallel to the green one as it was when she started (light red). Now it is oriented at 90 degrees to it, even though it remained pointed in the same direction for the entire trip!

Figure 8

Furthermore, when she gets her map out, she sees that unlike her brother, she has traversed an equilateral triangle whose inner angles add up to 270 degrees rather than 180 degrees. And even though the final leg of her journey was noticeably shorter than Freddy’s, she traversed a larger region. Having studied higher mathematics at Flatland University, she is familiar with higher-dimensional spaces than the two dimensional one she knows, and an equilateral triangle with three 90-degree interior angles sounds suspiciously like a higher-dimensional sphere. Sure enough, when she measures the area enclosed by her journey, she finds that it is given by,

A = \frac{\pi R^{2}}{2}           [Eqn. 3]

Where R is a parameter that behaves mathematically like the radius of a three-dimensional sphere even though in her universe, there is no third dimension to contain one.

Note that Cathy’s conclusions were based only on measurements of distance and area, and the orientation of a vector she carried with her around a closed two-dimensional path. At no time did she step “outside” of her space into a third dimension from which the radius of a 3-D sphere could be observed. What she measured is simply a parameter that behaves like one in area calculations. Of course, Figures 5-8 are shown in 3-D perspective for heuristic purposes, but beyond that, there is no need for Cathy to postulate any higher dimensions to explain what she sees. As far as she knows, in her universe only two dimensions exist. How could Cathy’s two-dimensional universe be “spherical” when the sphere of our experience is a three-dimensional shape?

Straight vs. Geodesic

To answer this question, let’s go back to the turn of the 3rd Century B.C. when the Greek Mathematician Euclid published his Elements. In it, he laid the foundation of geometry in our three-dimensional space and Freddy’s two-dimensional one with five axioms, or postulates. Of these, four are interdependent in that each one can be formally derived from the remaining three. The remaining one, his fifth postulate, he stated as follows,

If a line segment intersects two straight lines forming two interior angles on the same side that are less than two right angles, then the two lines, if extended indefinitely, meet on that side on which the angles sum to less than two right angles. – (Heath, 1956)

It follows from this that if the two interior angles formed are equal to two right angles, those lines will never meet. Though Euclid doesn’t specifically say so, this would make the two lines parallel, which led 19th Century Scottish mathematician John Playfair to restate it in what today is perhaps its most popular version,

There is at most one line that can be drawn parallel to another given one through an external point.

For centuries, the fifth postulate troubled mathematicians because reasonable as it may seem, it’s entirely ad hoc. It has no interdependence with the other four and is superfluous to a complete formalism of Euclidean geometry. It was only a matter of time until people began to wonder what geometric doors would be opened if it were discarded.

The first step in that direction is a reexamination what we mean by straight and parallel. Like Freddy, most of us think of a line as straight if it is one-dimensional in the sense of having no curvature—or more formally perhaps, if all points on it share a common tangent vector in one direction. Likewise, we think of two lines as parallel if they lie within a common two-dimensional plane and are aligned in the same direction with no intersection point. Indeed, this is how mathematicians defined both terms for many centuries, and to this day Euclid’s fifth postulate is often referred to as his parallel postulate. But if it proves to be superfluous to the formalism of Euclidean geometry then non-Euclidean geometry becomes possible and these definitions will need to be revisited.

Without the fifth postulate, on an N-dimensional manifold (or space), a curve connecting any two points A(x_1, x_2, ..., x_N) and B(x_1, x_2, ..., x_N) is said to be straight if, and only if it is the shortest distance between them on the manifold. In a flat space like Freddy’s (or ours), this reduces to our intuitive definition above, but that definition alone does not constrain manifolds to be flat. This suggests that if we want to quantify how paths between events are traversed in universes like Cathy’s, our mathematical descriptions need to be revised, and our parallel transport thought experiment gives us a clue as to how.

Modeling Curved Geometry

A full mathematical treatment of general relativity is beyond our scope today, but we can get our feet wet with an overview of the tools it will require. To model any N-dimensional space, be it flat or curved, there are two fundamental requirements we must meet.

First, we need a way to describe not only distances, but angles. To do that we will need to define at least two vectors at every point on it, r^{\mu} and r^{\nu}, where the indices \mu and \nu denote the N coordinates of each. Strictly speaking, they can be specified in any coordinate system of our choosing, and oriented in any non-parallel direction we like, but ideally, we want them to be orthogonal to each other (as shown in Figure 9) so that they define a coordinate system/s themselves. With these, we can then use a vector inner product, or dot product of them to define a matrix function  g_{\mu\nu} whose squared diagonal terms can be summed to give the squared distance along any interval, and whose off-diagonal terms are the dot product projections of each vector’s components onto those of the other. This function, which is referred to as the metric tensor,1 contains within its N^{2} components a description of all lengths and trigonometric relationships between the two vectors.2  [Aron discusses this at length in his 2012 post All points look the same.]

Figure 9

Neglecting time for simplicity (we’ll get to this later), in a flat 2-D space like Freddy’s, the two vectors will not have components that lie along each other so the off-diagonal terms will be zero, and the vectors are chosen so that their lengths define units in our chosen coordinate system,3  the diagonal terms will be 1 and the sum of their squares defines the Pythagorean theorem. Thus,

            [Eqn. 4]

Second, we need to ensure that our models preserve one of the most sacred principles in physics—namely, that the universe exists independent of us, so its behavior should be independent of how we choose to describe it. If the most fundamental laws of physics are different here and now in this coordinate system and units than it is there and then in those coordinates and units, that would imply that we have an unreasonably unique status in it. In our hearts, we know that isn’t the case, so our descriptions of it should look the same in all frames of reference and units. In physics this is referred to as the principle of general covariance.

To do this we need to account for the fact that some quantities behave differently under a change of scale in coordinate system units. For instance, if the vector r^{\mu} is one meter long, it will have a length of 1 in a coordinate system specified in meters. But if the scale is changed to centimeters, its length will be 100. The same will apply to angles. The vector itself remains the same—what has changed is its representation in a rescaled coordinate system. Quantities that behave this way are said to be contravariant because their size will vary counter to variations in the scale of units they’re represented with.

On the other hand, there are quantities such as gradients for which this isn’t the case. A 6% grade is a 6% grade whether we specify it in meters/meter or cm’s/cm, so rescaling coordinate systems will vary length specifications along any coordinate axis, but not the gradient in that direction. The metric tensor g_{\mu\nu} is such an object. As we’ve seen, it’s effectively a generalized dot product between local coordinate system axes. Since its components give their projections onto each other, it behaves like a gradient under coordinate system transformations. Quantities like this are said to be covariant because they retain their values regardless of how their coordinate system scale is varied. The difference is shown in Figure 10 (Wikimedia, 2018).

Figure 10

This may seem like hair-splitting, but when we move from the realm of absolute flat spaces to that of curved geometries, the difference matters. Some quantities like vectors, lend themselves to a contravariant description whereas others, like gradients, lend themselves to a covariant one. In the parlance of general relativity, it’s customary to specify the indices of the former with superscripts (“upstairs”) and the latter with subscripts (“downstairs”). Each type of tensor can be converted into the other by multiplying with an appropriately dimensioned factor (which is referred to as “raising or lowering indices”), but things are a lot clearer when we stick to representing each in the form that is most natural to them. As such, objects like vectors whose specifications vary under a rescaling multiple coordinate axes are typically specified with “upstairs” indices and those like the metric that behave more like gradients use “downstairs” ones.

With these qualifications, let’s revisit our earlier parallel transport experiments and put some flesh on the bones. In Figure 9 we saw that in Freddy’s universe, r^{\mu} and r^{\nu} will be the same everywhere and so will g_{\mu\nu}. It makes no difference where (or when) we place any coordinate system. But what about Cathy’s universe? At point A, a small surrounding region will be approximately flat and represented by a tangent plane containing r^{\mu}, r^{\nu} centered on it (Figure 11). Now, let’s define a third tangent vector s^{\sigma} along our parallel transport path from A to B.

Figure 11

Once again, we walk the path from A to B in the direction ds^{\sigma} as in Figures 5 and 6, carrying the tangent plane and r^{\mu} and r^{\nu} with us (Figure 12).

 

Figure 12

At each point in the path, r^{\mu}, r^{\nu}, and s^{\sigma} are still oriented in the same directions with respect to any local coordinate system, and the latter remains parallel to the path we’re travelling. When we arrive at B, we see that things still look the same to us as they did when we started. But this time the local tangent plane and coordinate systems we carried with us have twisted with respect to where they were at A and no longer looks the same to an observer who stayed behind.

In Freddy’s universe, one tangent plane uniquely spans the entire space. All distances and angles look the same from any reference frame within it, and carrying vectors such as r^{\mu} and r^{\nu} from one point to another is just a matter of summing displacements along any given path between them. But in a curved space like Cathy’s, we need a mathematical object that not only describes displacements along a path, but also one that maps that path onto the local tangent planeas it rolls across the curved surface as shown in Figure 13 (Wikimedia, 2023).

Figure 13

This object, which mathematicians refer to as an affine connection, allows us to describe vectors along any path through a larger curved space in terms of a fixed coordinate system within the local tangent plane at any point. An infinite number of such connections are possible but there is one, known as the Levi-Civita connection, that is a natural choice for spaces that have a well-defined metric tensor at every point because it allows us to define a derivative (or rate of change) along a curved space path that generalizes the usual mathematical rules of vector calculus in locally flat tangent plane regions to the larger curved space. This covariant derivative (which we denote with the nabla symbol 4) will need to have two parts and is given by,

\nabla_{\mu} = \partial_{\mu} + \Gamma^{\sigma}_{\mu\nu}           [Eqn. 5]

For an infinitesimal displacement along any path, the first term on the right is the gradient with respect to the local tangent plane as defined in the usual flat space manner. The second term is the rate at which the tangent plane itself (and the covariant metric tensor embedded in it) is changing in the direction of a contravariant displacement ds^{\sigma} in the direction of a tangent vector to the path. As such, it will be matrix function with three indices, two of which are best represented as covariant and a third contravariant one which we will denote with the index \sigma. This function, which per convention we designate with a capital Greek Gamma, is known as a Christoffel symbol. Since it requires three indices to fully capture the evolution of the metric tensor, in Cathy’s space it will have 23, or 8 components to her metric tensor’s 4. We refer to Christoffels as “symbols” because they aren’t true tensors in that they aren’t globally frame-independent until multiplied by an infinitesimal displacement in at least one direction. And as shown, equation 5 doesn’t make sense because the indices on the right and left sides don’t agree with each other. More properly, it defines a mathematical operator that must act on something to produce a meaningful equation. Applying it to ds^{\sigma} gives,

\nabla_{\mu}s^{\sigma} = \partial_{\mu}s^{\sigma} + \Gamma^{\sigma}_{\mu\nu}s^{\nu}           [Eqn. 6]

With the upstairs and downstairs \nu in the second term cancelling, this equation is now consistent across indices and the Christoffel term behaves like a tensor. This path derivative will look the same from every coordinate system in Cathy’s curved space. In flat spaces like Freddy’s, the tangent plane is the same everywhere and unchanging so the Christoffel term will vanish leaving us with the usual Euclidean directional derivative we learned in first-year vector calculus.

 

For today’s purposes we needn’t worry about how these equations were derived. The important thing is to understand why curved spaces require these kinds of mathematical tools rather than the familiar ones of Euclidean geometry, and how they reflect curvature in multiple dimensions without additional dimensions to “curve into.” If you’re like me, the latter point is the biggest stumbling block. It’s one thing to know that curved spaces are mathematically possible without additional background dimensions. But it’s another thing altogether for three-dimensional Euclidean space beings to visualize them. Space (or spacetime) can be curved in one of two ways: positive, or negative.5 Positively curved space is spherical and, if extended far enough, finite and closed. In our previous example, Cathy’s universe is a spherical one. And as we saw, the interior angles of a triangle in such a space add to greater than 180 degrees. Her space is finite in size, and travelling in a straight line in any direction will eventually return you to where you started from. Negatively curved space is saddle-shaped and has hyperbolic geometry. The interior angles of a triangle in it would add to less than 180 degrees, and like flat Euclidean space, it extends to infinity in all directions. Figure 14 shows both as compared to flat space.

Figure 14

It’s easy to visualize two-dimensional curved spaces like these in isometric views that show their contours in an additional dimension. But what would they look like where there was none?

In the case of a positively curved space, we can’t do this because there is no way to represent a path that returns to where it started in the same number of dimensions.6 But for negatively curved spaces that extend to infinity, we have a visual example in the art of 20th Century Dutch graphic artist M.C. Escher. Among other things, Escher was known for artistic renderings of mathematical concepts including symmetries and tessellation. His Circle Limit collection of wood carvings depict repeating image patterns whose changing shapes from the center outward are a tessellation of hyperbolic geometry on a disc into right triangles. His 1959 work Circle Limit III (Figure 15), widely regarded as the best in the series, does this with patterns of fish.

 

Figure 15

There are many ways to tessellate geometric spaces and none are perfect, including this one. But if Cathy’s two-dimensional space was negatively rather than positively curved, this would be a reasonable representation of how it would look to her. If she walked a parallel transport path through it as in figures 5-8 taking the size and orientation of the fish as indicative of distances and angles, upon returning to where she started, she would find that the distances and interior angles she traced would be like those in the negatively curved saddle in figure 14. And if she travelled a straight geodesic path in any direction indefinitely, she would asymptotically reach infinity as she approached the rim. The disc is two-dimensional, but the geometry embedded in it behaves as though it were a saddle-shaped sheet in three dimensions even though the third dimension isn’t there. The underlying mathematics of its hyperbolic (saddle) geometry are embodied in Equation 6. And while we have until now restricted ourselves to two-dimensional spaces for ease of illustration, notice that the indices in its terms can assume any number of values, not just two. As such, it generalizes to any number of curved dimensions, none of which need any “higher” dimension/s to curve into.

There is, however, one dimension that we’ve conspicuously ignored until now… time. We live in a universe where not only length, breadth, and height can be curved, but duration can be as well, and curved spacetime ups the ante in several important respects that we’ll dive into in Part II. So, stay tuned!
 
Curvature II: Spacetime
 

Footnotes

1)   In mathematics, tensors are matrix functions that define a multilinear relationship between sets of objects in a vector space that preserve their identity in any coordinate system or transformation. Vectors can be thought of as a one-dimensional tensor (that is, a tensor with only one column or row). The dimensionality of a tensor’s matrix array (as specified in the number of indices it requires) is referred to as its rank R, and the number of components it will have in an N-dimensional space is given by N^{R}. Thus, g_{\mu\nu} is a rank 2 tensor that in Freddy’s 2-D space will have four components, and in our 4-dimensional spacetime has 16.

2)   Strictly speaking, the metric tensor isn’t really a true dot product. Rather, it is a generalization of the familiar dot product of Euclidean geometry to the pseudo-Riemannian geometry constrained by special relativity, where time behaves differently than space (more on this in Part II). But for our current exploration of 2-D spatial curvature, this needn’t concern us.

3)   Mathematicians refer to this as an orthonormal basis that spans the space.

4)   In mathematics, the nabla symbol (\nabla_{\mu}) is known as the Laplace operator. It is a shorthand reference for the gradient (first derivative) in the direction of a vector defining the \mu coordinate system; That is, \nabla_{\mu} = \frac{\partial }{\partial x_0} + \frac{\partial }{\partial x_1} + \frac{\partial }{\partial x_2} + \frac{\partial }{\partial x_3} where the index \mu = 0, 1, 2, 3. This representation of a gradient in a particular direction is also referred to as the vector’s divergence.

5)   The reasons for this are mathematical and beyond the scope of this discussion.

6)   This is because spherically curved space has a different topology than flat and negatively curved spaces. In mathematics, topology is the study of a manifold’s geometric properties that are preserved when it is stretched or deformed without cutting or sewing, opening or closing holes, or passing it through itself. Negatively curved space has the same topology as flat space because a flat rubber sheet can be stretched to form a saddle. By contrast, a positively curved space cannot be flattened or deformed into a saddle without cutting and forming edges (e.g. – a Mercator projection). There is no way to create a flat representation of it that preserves great circle paths that end where they began without encountering an edge. Likewise, a toroid (donut) cannot be deformed into a sphere or a saddle without cutting and sewing edges, so it has a higher-level topology than negatively or positively curved spaces.

 

References

Heath, T.L. ed., 1956. The thirteen books of Euclid's Elements. Courier Corporation. Online at https://books.google.com/books?hl=en&lr=&id=mvBIAwAAQBAJ&oi=fnd&pg=PP1&dq=euclid+elements&ots=ed2L7zetPz&sig=wPKfMQ22SZvf4gF_83USfDwb0oY#v=onepage&q=euclid%20elements&f=false. Accessed Sept. 28, 2023.

Wikimedia. 2018. Image courtesy of Jacob Bertolotti. Online at https://commons.wikimedia.org/wiki/File:Covariantcomponents.gif. Accessed Sept. 28, 2023.

Wikimedia. 2023. Image courtesy of Silly rabbit, CC BY-SA 3.0. Online at https://commons.wikimedia.org/w/index.php?curid=2615879. Accessed Sept. 28, 2023.

Posted in Physics | 2 Comments

Book by St. Tom Rudelius (and me, a bit)

So my friend St. Tom Rudelius is a physicist who works on string theory, QFT, and early universe cosmology (e.g. the theory of inflation).  He is also a brother in Christ who I have had the privilege to both mentor, and learn from.

He has just written a book about his conversion to Christ (it's a pretty interesting story, involving rather more "polygraph tests" than this sort of story usually involves) and also his experiences as a Christian in academia.  The book, which was just released today, is called:

 

 

 

 

 

 

 

 

 

 

 

I was asked by the publisher to include an excerpt from the book to help promote it.  Completely disregarding their proposed selections, I have chosen one of the later chapters of the book, after he's already become a Christian:

People often ask me what it’s like to be a person of faith in the field of science. It’s a hard question to answer, because my experiences have varied widely.

Sometimes, physicists will ridicule religion. Once, while visiting the University of Texas to give a talk on my research, I went to lunch with a number of physicists, including the late Nobel laureate (and outspoken atheist) Steven Weinberg. Unaware of my religious leanings, Weinberg began the lunch with a pointed question toward the antievolution movement: “Do all these people who reject evolution also reject cosmology?”

I thought about explaining the difference between young earth creationists and old earth creationists, but ultimately held my tongue.

Sometimes, physicists simply steer clear of religious topics. One day when I was a postdoc at the Institute for Advanced Study in Princeton, the man whose donations to the Institute helped pay my salary came to have lunch with Ed Witten—quite possibly the greatest living theoretical physicist, if not the smartest man on earth—and me. A quick online search had made the donor aware of my religious views, so he spent the entire lunch asking me (very respectfully) about my opinions on religion and politics. It was probably the most stressful conversation I’ve ever had—talking about Jesus and Donald Trump with the smartest man alive and the man who paid my salary.

During the entire conversation, Ed Witten was surprisingly quiet. His only remark came when we were discussing God’s miraculous intervention. “I think a lot of people wish God would intervene more often,” he said.

Sometimes, physicists respect religion. Several of my colleagues have expressed admiration for my religious faith, or religious faith in general, though they themselves do not have any religious convictions.

Sometimes, physicists embrace religion. I don’t know very many Christians in my field, but whenever I meet one, I feel an immediate kinship. Our scientific drive for knowledge pushes us to learn as much as we can about the physical universe, and as Christians that same drive pushes us to learn as much as we can about God. The result is a common language of science, theology, and philosophy not so different from the “twin telepathy” my brother and I have shared since childhood. Though sometimes it is discouraging that so few of my colleagues embrace religious faith, it is encouraging—perhaps even more so—that the ones who do are so strong in their faith and so capable of defending it intellectually.

In much of the world, there is intense animosity, and sometimes even violence, between people of differing religious faiths. Perhaps it’s because we religious physicists represent a minority in our world, but I’ve certainly never felt anything like that from my Jewish, Muslim, and Hindu colleagues. And I hope they’ve never felt anything like that from me. Rather, there seems to be a sense of solidarity among religious scientists. Though there are important differences between our faiths, there’s an even deeper sense of mutual respect among us: I’ve probably received more comments of admiration regarding my faith from Jewish colleagues than I have from Christian ones, and a Muslim colleague once told me that my public interviews and articles on science and God had strengthened his own faith.

On the whole, though, I can say with certainty that I have never felt persecuted or personally attacked for my faith. There are places in the world where Christians are suffering for their faith. But America is not one of those places. I can go to church, pray, read my Bible, and even write books like this one without fear of losing my job. Some of my colleagues may not agree with my faith, but fortunately my success in physics depends on my ability to do physics, not on how I worship in my free time.

Though science and faith are often viewed as enemies, I can also say I have felt less hostility toward religious faith in the upper echelons of physics than at the lower levels, or in the soft sciences or humanities. Anthropology, history, and religious studies departments are famously dismissive of Christianity—a trend many of my Christian friends and I experienced during the course of our university studies.

One of my friends who studied chemistry at Princeton had a high school science teacher who forced the class to learn the definition of a so-called scientific theory—an explanation for some natural phenomenon supported by a vast body of evidence—to refute the common creationist retort that “evolution is only a theory.” But when he got to college, my friend soon realized that such definitions are nonsense: In practice, scientists use the term theory to describe many different things. Some theories, like quantum field theory, are among the best tested phenomena in all of science. Other theories, like string theory, lack any experimental verification whatsoever.

My high school physics teacher—who was one the best and most important teachers I ever had—occasionally made snide remarks about religion. Yet at Cornell, Harvard, and Princeton, I met several religious physics professors. One professor even suggested to his class that God might be the best explanation after all for the fine-tuning of the universe for intelligent life—and he wasn’t even a theist.

Now, it’s also true that most of my extraordinarily brilliant colleagues do not embrace religion. But I’ve found that their reasons are generally quite ordinary. If you ask the average atheist why he or she doesn’t believe in God, you’ll probably get some version of the problem of evil: “If God is all-powerful, all-knowing, and all-loving, why do evil and suffering exist?” If you ask one of the world’s most brilliant scientists why they don’t believe in God, you’ll probably hear the exact same thing.

That’s not to say that the problems of evil and suffering are easy for theists to deal with. It’s simply that the most brilliant minds don’t have a huge advantage over others when it comes to questions of faith. We all have basically the same questions, objections, and doubts. In my experience, the ones who find answers to these questions are typically those who need answers the most. Personally, before Steve’s conversion and subsequent conversations with me, I never felt much need for religion, as I was generally able to get by on my intelligence alone. Perhaps other scientists feel similarly.

Finally, I have found that most scientists—even nonreligious ones—believe in some sort of power greater than ourselves. It’s very common to hear physicists refer to Nature as a sort of placeholder god. For example, Ed Witten once said in an interview, “If I knew how Nature has done supersymmetry breaking, then I could tell you why humans had such trouble figuring it out.” There is a widespread acknowledgment that Nature has chosen a particular way for our universe to be, and it could have chosen something different.

What’s the difference between this Nature and the God (capital G) I believe in? I think the biggest difference is simply that Nature doesn’t really care much about the affairs of humanity, whereas God does. Most everyone would agree that Nature has a preference for order, simplicity, and beauty, but many balk at the suggestion that it would concern itself with the affairs of one particular species on one little insignificant planet. We humans are, to quote astronomer Carl Sagan, nothing but “a mote of dust in the morning sky.” [1] Why would God care about us?  

[1] Carl Sagan, Cosmos (New York: Random House, 1980).

To this, I like to point out that size is not a very good measure of value. I care more about the life of a baby than I do about most galaxies. I care more about the ten-nanometer transistors that make my computer work than I do about distant stars. And even as someone who studies black holes and the big bang for a living, I find nothing more incredible about the cosmos than the fact that it somehow birthed intelligent, conscious beings like us.

Ultimately, one can choose to view the size of our universe as a sign of our insignificance, or one can choose to view it as a sign of the great significance of its creator—a creator whose attention is not divided, who built and sustains the intricate workings of the cosmos, yet who simultaneously cares enough about humanity to become a human himself, to experience pain, suffering, and death so that we could have life.

Perhaps you noticed that my name is also on the front of the book, in much tinier yellow letters at the bottom.  (Or more likely, you didn't and are even now scrolling back to see if my claim is true.)  This is because I was asked by St. Tom to write a foreword to his book.  (And not only that, I did.)  My foreword begins as follows:

Foreword

(from this formative experience with the publishing world, I have learned that the word has an "e" in it) but after that it goes on to say:

The book you are holding is a remarkable one. There are lots of books out there promoting Christianity, by a type of person you might call salesmen. The goal of a salesman is to produce a watertight and squeaky-clean argument, to convince you that only one position is intellectually respectable, and fully capable of servicing your needs. He is afraid to admit any weakness in his arguments. He is afraid that if he talks honestly about his own doubts and struggles, his audience will take it as a reason to reject the product he is promoting. If you want a book like that, I suggest you look elsewhere. My friend Tom is not a salesman. But he is a person who cares deeply about what is real, both in scientific and religious contexts. And because of this, he is also unafraid to share his spiritual doubts and struggles, both before and after he became convinced that Christianity is objectively true.

After that, the foreword includes eleven more juicy paragraphs, and importantly the only way to read them (if you don't know about libraries) is by buying the book.  You can do this by clicking on one of the following links:

Amazon

ChristianBook

Tyndale

Target

That's right, you can now buy the equivalent of one of my blog posts, at the same store you can get detergent and kid's T-shirts from!  But, you should probably also buy the book to read an interesting and sincere account from Tom, about his obstacles coming to Christ and his emotional struggles with faith afterwards.

Now, you are going to buy the book at any time in the future, it would probably be helpful to Tom if you would buy it ASAP, for example TODAY, so it can go into the early sales figures that make the industry decide whether this book is hot stuff or not.  Sorry, I don't make the rules of worldly success in the publishing industry, that's just how it goes.

Having said that, some of you may be tempted to write comments asking, well when are you (Aron Wall, PhD) going to write your own book about Sciencey-and-Religiony stuff, and not just a foreword or backewards glued onto somebody else's book?

Well, as you can probably tell from my recent blog performance: I'm just way too busy (with mentoring PhD students and postdocs, parenting my 2 & 4 year olds, quantizing gravity, and doing faculty busywork) to get any useful writing done, for the most part.  Nevertheless, you should expect some book about the Fine Tuning Argument for God and/or the Multiverse to appear under my name (as well as that of my coauthors, philosophers John Hawthorne and Yoaav Isaacs), some time in the next oh 1-50 years from now.  Just thought I'd give you a heads-up about that since a file looking deceptively like a rough draft basically already exists, more or less.  Mostly less.

(If any skeptical promotion committees are reading this post, I promise I spent very few of my months working on the Fine Tuning book, and any deficit of actual physics papers is explained by the other stuff in my life...)

Anyway, life is short so don't save your money for a book that might or might not come out in the next couple of years.  Instead BUY TOM'S BOOK NOW (if you feel led to do that) and trust that you'll have the spare change to buy mine later.

[Disclaimer: I understand that I will be receiving a free copy of Tom's book in the mail.  But it will come too late to change my opinion of the book—I will always think of Tom's manuscript primarily as a Word file.  I'm sure the publishers put a lot of effort into making it look like a real book; but I'm sorry, that's just the way it is.]

Posted in Links, Reviews | 15 Comments

Followup on the Moral Argument for Theism

A commenter named Nikki argued against my post Fundamental Reality XII: The Good, and the Not.

Nikki writes:

I don't think the post's argument works - I'd argue that non-theistic morality can be objective and well-grounded, or at least be no worse off in those regards than theistic morality is.

So the first part of this post that really jumped out at me is the claim that if morality is objective, it must be like a mind. Frankly, to me this seems not only false, but a category error. Morality is things like systems, principles, rules, etc. - I'm not sure what the exact best word choice is. The point, though, is it is a thing that minds use, but not in and of itself a mind. You describe morality as approving or disapproving certain things, but this seems to be conflating things like "this abstract system contains claims that X is good/bad," which could validly be said about morality, and "this abstract system itself consciously judges that X is good/bad," which could not. It is us who use morality to consciously make those judgements.

As an analogy, personality traits are part of minds, but not minds themselves - to speak of them, by themselves, being conscious, thinking, willing, etc. would be a fundamental mistake. (Though Inside Out was a pretty fun movie). I'll admit though, I don't actually think that's the best analogy. I'd argue the set of laws of logic or mathematics are an even better example of something that is a feature of minds - but is not, and could not possibly be, a mind in itself. However, you've said in an above comment that logic is also a description of God's character.

(Perhaps a bit of a sidetrack here, but I don't think this could be true either. I believe that you've stated elsewhere that while you believe God is metaphysically necessary, he is not logically necessary - but of course, it is logically necessary that the laws of logic or mathematics are true. I don't think the dependence you're arguing for could work, even if God exists in some sense. That said, as one might guess, I don't think God is metaphysically necessary in the first place.

In fact, I have doubts that there is even a "metaphysical necessity" distinct from logical necessity at all. I find Chalmers' arguments in his paper "Does Conceivability Entail Possibility?" fairly convincing in this regard. I do think there are some weak points, but it seems to me that at least it shows that even if there is a metaphysical modality separate from logical modality, we don't currently have a good reason to believe in it. I know there are several relevant arguments on this blog, but well, I can't discuss every single reason for and against the existence of God in this post, so here I'm trying to stick with things related to the original topic/what's been mentioned in previous comments on it. I might debate the other arguments later.

As a note, Chalmers' arguments there are important for the case he makes that consciousness is not physical, because they counter the reply of some materialists that consciousness is metaphysically the same as a physical property, even if it cannot logically be derived from other physical facts. Others have argued that this causes problems for theists who both defend the metaphysical necessity of God and the non-physicality of consciousness. I suppose this may not apply to you because you've said you can't rule out that consciousness is physical in some sense, as in what Chalmers calls "Type-B Materialism," but I did think it was interesting).

Alright, back to the main topic. Does an objective morality depend on God? The whole field of moral philosophy is certainly not something I can fully describe in one post, but I'll start with something interesting you said in your own previous post in this series:
"Even people who say there's no such thing as ethical truth suddenly sound quite different when somebody treats them unfairly."

I suspect that in that statement is at least a hint at what the basis for a nontheistic objective morality might be like. If there is an objective morality, I think it has something to do with the symmetry between you and others - if you don't treat others well, what's to prevent them from doing things to you that you don't want? Even if evil may sometimes have short-term rewards, people committing acts like theft or murder or terrorism ultimately make things worse for everyone, including themselves. Note that these statements do not depend on God to make them true. And I think several strands of thought, such the Golden Rule, Kantian morality, Rawls' veil of ignorance, and even some game-theoretic analyses, among others, all point towards something like this in a sense.

Now, this may not be very compelling - I'm being vague and have not spelled out a fully detailed nontheistic system. Furthermore, many of the systems I've cited actually contradict each other. Nevertheless, I think that there are important shared elements that don't depend on a belief in God to be convincing (well, Kant's morality was theistic and the Golden Rule is a part of many religions, but I don't think everything along the lines that I've mentioned is). So it seems that the claim that no secular account of morality can possibly succeed isn't very certain. I'll note that you linked in your previous article to the SEP's article on Moral Naturalism, but merely said those systems were "problematic" without really discussing the individual ideas presented there, although there are many important nonreligious thinkers whose ideas on morality are much more detailed than mine. (I won't complain about that too much though - after all, I'm not discussing every form of theistic morality in this post myself).

Some more notes: 1. Speaking of moral naturalism, even on atheism, that isn't the only option available for an objective morality. While I agree naturalism and atheism are often found together in practice, it is still possible for an atheist to be a non-naturalist, including about morality. So even if morality cannot be justified on naturalism, you would have to show that God specifically is the only one who can ground morality, not some other non-natural element.

2. Above, Scott Church argues that on naturalism, the universe does not care about us and we are fundamentally unimportant, so it cannot ground objective morality. But the universe itself does not have to care about us/be a moral agent for morality to be objective! I'd argue that if morality, say, applies to all rational beings, it is objective, and the universe not obeying it does not matter because the universe is not a rational agent. The laws of rationality themselves are a good analogy for this - the universe, itself, does not reason, and it requires minds to use reason, yet the standards of rationality are fully objective (and not derivable from physical equations, by the way). And even on theism, it is agreed that some things, like inanimate objects, are not and cannot be moral, yet again, that does not prevent morality from being objective. Related, while pure pleasure-maximization/pain-minimization has several well-known problems, so I doubt that's the full objective morality, I do think there are non-arbitrary reasons why those are at least important. They are necessarily important to us by their very nature - no one can truly be indifferent to them even if they claim to be. And even if the universe does not care about them, I take the anti-nihilistic view that it is precisely the fact we care that matters - it's not as if the universe has any rule against that!

3. I've seen this part stated before in some other comments on the blog, but I think it's important enough that I'll state it again (especially since unless I'm missing it, I don't think I've seen a response). Escaping the Euthyphro dilemma by saying that God is identical to goodness can only work if we have good reasons to believe that the two could possibly be identical. I don't think we have those (unlike for the triangle case, in which we do have reasons to believe that "having three sides" and "having three angles" are the same, even though those are logically necessary), but we do, in fact, have reasons to believe the opposite. As I wrote at the beginning of my post, if God is to be viewed as even like a mind, he cannot possibly be identical to morality even if he is an (ultimately) moral agent. For instance, one of the important reasons to consider God like a mind is that he is supposed to be able to take actions, but morality cannot, by itself, take actions. (Also, I'll admit I don't know whether your analysis of Plato is accurate, but even if it is, it's generally fine to take inspiration from an argument and adapt it to your own views. After all, in the original article, you said you used "Hume's Is-Ought dictum in a manner which he would have thoroughly disapproved of!")

As a final statement, I don't think theism is actually better at convincing people of being moral than secularism. There's some evidence that nonreligious people are even more moral than very religious people, but interpretations are controversial and I'm focusing more on purely philosophical points here. (I do suspect nonreligious people being more moral than the religious, if true, would be a particularly big problem for theism and theistic morality. I think the evidence at least shows that the nonreligious are generally not less moral than the religious, but you've agreed in another article that for some senses of "good," religion is not strictly necessary for it, so that may not be a big problem for you). But anyway, you've agreed that not all rational people might be convinced by theistic arguments, and it's been pointed out above that you can always ask questions like "Why should you follow God's commands?" so that seems to be an issue. Of course, you might very well always be able to ask similar questions about any nontheistic system, and rational people might not find it convincing. But my point was that secular morality is at least equal to theistic morality in this regard, and while this is a bit speculative, perhaps some of the reasons above might make the former even more convincing than the latter.

My reply got pretty long, so I'm turning it into a blog post.

Dear Nikki,
Welcome to my blog, and thanks for your interesting comment. However, I am not sure that your arguments are actually directed against the specific argument I am making. Here are some replies (not in the order of your points):

I. Objective Morality is a Premise in the Moral Argument

You make a good case defending this proposition: It is possible for a non-theist to rationally come to believe in the existence of an objective ethical system, without thereby coming to believe in God. However, I also believe that this is the case!

In fact, if this were not true, there would be little rhetorical point in presenting a Moral Argument for God's existence.  In order for an argument for God's existence to be capable of being convincing, there have to be some people out there who agree with the premises of the argument, but have not yet realized that the conclusion follows (or at least, is made more probable) by the premises.  I obviously do not deny the existence of non-theistic moral realists, because they are the target audience for my post!  (That is why I presented an argument for ethical realism in part XI before describing how  I think Theism grounds ethics in part XII.)

Now obviously, if the a nontheistic argument for objective ethics happened to take the form of an entirely satisfactory reduction of concepts like ethical obligation into naturalistically acceptable terms—e.g. in terms of physical facts of the sort that even Sean Carroll would accept—then the Moral Arguments for Theism would fail, since there would be no additional work for God to do in terms of grounding ethics.  (There might still be a need to ground the laws of physics in some way, but no additional and separate need to ground ethical truths.)  But of course, if you could show that this were true, you would have just solved a very famous and important problem in philosophy!  So I sort of doubt you really think that we can know this to be the case.  And if we cannot know it to be the case, then there is room for discussing non-naturalistic groundings of ethics, in a probabilistic argument for Theism.

You sketch some ways in which you think an non-theistic grounding for objective ethics might work (which fall into the rough family category of what I called "Kantian approaches'' to ethics in part X).  As I explicitly stated in that post, Kantianism is not as friendly to the Moral Argument, as Platonism or Aristotelianism is; although I don't think it is utterly hopeless on that front.  (Kant himself made a sort of pragmatic argument for Theism from Morality, but he didn't agree with metaphysical arguments of the sort I'm discussing.)   The only conclusion I explicitly drew from Kantianism was:

If Ethics can be deduced rationally as in the Kantian system, then one can at least deduce that if the Universe originates from something like a mind, that mind should also be able to appreciate ethical truths.

So the point you are making was to some extent already acknowledged in this series.  (Of course, on classical forms of Theism, where God is something like the ultimate Reason or Logos behind the Universe, this would still end up identifying God with moral goodness in some deep sense; but such classical views are necessarily bordering on Platonism anyways...)

B. Moral Naturalism and Non-Naturalism

By the way, I revisited the SEP article, and found to my dismay that it had been edited in a way that removed (without refutation) some of the critiques of Moral Naturalist positions. Here is the original version of the article.  If you look, for example, at the original article's section 4.3, you can see what appears to me to be a pretty desperate attempt by Jackson to make naturalistic ethics work, together with (what appears to me to be) a pretty strong refutation in terms of the permutation problem.  But the main point is not the refutation of that particular idea, but that I don't see any way forward mentioned in the article which doesn't seem to have serious problems.

You write:

Speaking of moral naturalism, even on atheism, that isn't the only option available for an objective morality. While I agree naturalism and atheism are often found together in practice, it is still possible for an atheist to be a non-naturalist, including about morality.

Yes, obviously.  Such views exist (which is why I mentioned them in part X of this series). In fact, individuals with such views (e.g. Moral Platonists) are closer to being the target audience of this post, then perhaps you are.

So even if morality cannot be justified on naturalism, you would have to show that God specifically is the only one who can ground morality, not some other non-natural element.

No, because as I tried to make it clear at the beginning of this series that I wasn't trying to present a deductive, logically watertight argument for Theism.  As I said in Part I:

Even if there are no strictly deductive arguments (from indisputable premises), there are still going to be plausibility arguments pointing in various directions.  It's irrational to put too much faith in plausibility arguments, but it's also irrational to be completely insensible to them.

So the mere existence of logically possible positions, besides the one I argue for, doesn't bother me.  The question is which positions are most credible.

On the plausibility front, it seems to me that once you start modifying your metaphysics in order to accommodate objective ethics, it would be irrational not to take that into account when assessing the probability of other metaphysical hypotheses.  Ethical Monotheism is, among other things, the belief that a fundamentally good being exists.  The plausibility of this statement depends in part on what we think moral goodness is.  For example, on the view that:

1. "Morality is a emergent and subjective set of feelings found in some of the higher apes, conducive to their evolutionary survival, but having no basis in any metaphysical reality"

then the idea that there exists a fundamentally good being outside the physical universe—which did not evolve—is totally absurd.  On the other hand, if:

2. "moral facts are necessary truths, which tell us something substantive about the structure of non-physical realities",

then the idea of a fundamentally good being is, though not logically compulsory, at the very least far more plausible than on viewpoint (1) than (2).  Do you agree with that?  If so, then you are necessarily agreeing with me that the Moral Argument for Theism has significant probabilistic force.

[Notes: I am not saying these are the only possible views.  Also, hypothesis (2) does not necessarily deny biological evolution, as it is possible for evolved systems to recognize necessary truths such as mathematical theorems.]

C. The Role of Analogies

Let me remind you a bit of the context of my argument in the Fundamental Reality series.  In parts II-VI, I argued that it is plausible that there exists some fundamental reality which explains everything else, I discussed some properties this entity should have, and after reviewing various candidates I suggested that (based on the mathematical character of the laws of physics) the two most plausible metaphors for understanding this fundamental reality are:

* something like an equation
* something like a mathematician

Now it is important to remember that both of these ideas involve metaphors!  Obviously, if a Naturalist says that some equation provides the deepest truth about the Universe, that doesn't mean this assertion is being made about a set of chalk lines on a blackboard.

Similarly, if a Theist says that God is like a mind, that doesn't mean that this Mind is like our mind in every respect.  In particular, Classical Theism proposes a mind for whom there is no distinction between its subjective beliefs and objective reality, and also no distinction between its subjective preferences and objective morality.  This is obviously very different from evolved primate minds like our own!

You wrote:

So the first part of this post that really jumped out at me is the claim that if morality is objective, it must be like a mind. Frankly, to me this seems not only false, but a category error. Morality is things like systems, principles, rules, etc. - I'm not sure what the exact best word choice is. The point, though, is it is a thing that minds use, but not in and of itself a mind. You describe morality as approving or disapproving certain things, but this seems to be conflating things like "this abstract system contains claims that X is good/bad," which could validly be said about morality, and "this abstract system itself consciously judges that X is good/bad," which could not. It is us who use morality to consciously make those judgements.

and

As I wrote at the beginning of my post, if God is to be viewed as even like a mind, he cannot possibly be identical to morality even if he is an (ultimately) moral agent. For instance, one of the important reasons to consider God like a mind is that he is supposed to be able to take actions, but morality cannot, by itself, take actions.

I think perhaps you missed the amount of qualifying words I put into my reasoning.  What I wrote was (emphasis added):

But now observe that morality is at least a little bit like a mind, insofar as it approves or favors certain things, and disapproves or disfavors other things. So a fundamental morality would have something analogous to will or desire, and in that respect it would be more like a mind than like an equation, as in Theism.

The point here is not that an objective morality is exactly like a mind, but that it in certain respects more similar to a mind than (say) the equations of the Standard Model are, namely that the Standard Model does not encode any judgements that certain states of affairs are desirable or undesirable (as opposed to probable vs. improbable).

Now, obviously, when we say that God is personal, and can do things like forgive or create, we are adding more to our concept of God then is implied by the mere abstract notion of a metaphysical objective morality.  In my understanding of God, we are adding more to our idea of divinity than the idea of a Platonic form of the Good, but we are not necessarily taking anything away.

In other words, in my conception of God, God is such that he is good, not in an accidental (happenstance) way, but in an essential way, because all goodness in the universe in some sense participates in his goodness, just as all existence participates in his existence.  (The latter claim, of course, obtains for any fundamental reality which is taken to explain all other things.)

D. God Transcends the Abstract/Concrete Divide

Another commenter, St. David Madison, replied to your comment by saying (in part):

"You draw an analogy between morality and personality traits and then point out that personality traits are not conscious and do not themselves think. However, personality traits cannot exist without a personality that possesses those traits."

This is certainly a reasonable distinction to draw in general; and we could indeed escape from the supposed category error by simply replacing the words "objective ethics" with "that which grounds objective ethics, whatever it is."  But I think I am instead going to double down on this idea, and say that this supposed category distinction between abstractions and concrete objects breaks down when one is speaking about divinity, just as the distinction between particles and waves breaks down at the subatomic scale.  If God is the source of all else that exists, he must unify within himself the perfections of both abstractions (necessary, eternal, unchanging) and concrete realities (which are causally active, definite, individual etc).

This is indeed, already implied by certain sorts of religious language, in which God is portrayed not as some good or beautiful thing, but as the Supreme Goodness or Truth or Beauty or Life etc.  For example, in the Gospel of John, Jesus asserts his divinity by saying that he is the Way, the Truth, and the Life, which is not the sort of thing that a Positivist philosopher would consider a well-formed statement (a person cannot be an abstract quality).  But I am not convinced we can restrict our language in the way the Positivists wanted to do (I don't think Positivism even satisfies its own criteria of meaningfulness).  What this religious language points to, is an insight into the nature of divinity as a necessary being, in which all other realities are grounded.  A proposition about a created being can be true, but only the ultimate reality can be the Truth.  In other words, denying the applicability of the concrete/abstract distinction is not something I am doing merely to avoid a logical puzzle, but is already implied by standard religious language about God.

This sort of language about God makes Classical Theism radically different from traditional forms of polytheism, in which the gods are simply regarded as more powerful individuals than us, who still can be born/killed, have conflicts with each other, make mistakes etc.  Yeah, obviously the preferences of finite beings like ourselves can't possibly ground objective ethics, which was the whole reason why Plato went in a platonic direction instead.

Furthermore, I don't think we can avoid postulating this sort of concrete/abstract unification, simply by rejecting Classical Theism, as Naturalism seems to me to imply exactly the same thing.  For example, if the fundamental reality is something like a mathematical equation, then we are asserting that it is both an abstract piece of mathematics—which can in principle be understood by humans—AND ALSO the governing principle controlling the universe.  In other words, when a Naturalist does physics, they are still are postulating that the fundamental reality is a λογος, i.e. a rational principle.

Of course, I'm not saying that the equations we write on the blackboard, or in our minds, are strictly identical to the actual laws of physics, which obviously exist whether or not we ever discover them.  But if we asked, "what are the fundamental laws of physics like" we can't point to anything other than to our abstract human formulation of the equations, and then lamely add "except that it also exists as an actual concrete reality, in a way which transcends our human abstractions".

In the same way, objective morality exists even apart from human processes to reason about what is or is not moral—So I'm not saying, that this latter, social process of reasoning is equal to God.  Rather it is goodness as it actually exists (which our human reasoning is a mere approximation of) that is rooted in God's nature, as the ultimate Goodness that other things participate in.

E. Implications for Euthyphro

Escaping the Euthyphro dilemma by saying that God is identical to goodness can only work if we have good reasons to believe that the two could possibly be identical.

This is a strange way to discuss this subject, given that the (modern) Euthyphro dilemma is typically phrased, not in the form of a deductive argument, but in the form of a challenge to Theists to explain their beliefs more clearly.  It's phrased in the form: "Do you believe A, or B?" (both of which have unpalatable consequences).  But if A and B are not, in fact, exhaustive possibilities, because some other option C is conceivable—and if in fact C was the belief of most ethical monotheists historically, as well as myself—then merely pointing this out is sufficient to defuse the dilemma.

That being said, there is a good reason to think that, if God exists at all, he can ground morality.  Recall that God is, by definition, the explanation for all entities other than himself.  (That's the whole point of Mono-theism, to have only one ultimate entity.)  So if God exists at all, he either grounds or creates all other realities.  Now if there is objective ethics, then ethics counts as one of these realities.  Since it doesn't make sense to create ethics (since at least some ethical principles are non-arbitrary, necessary truths) then he must ground it.  (The same argument would hold for logic or mathematics.)

Now, to be clear, this is an argument that God grounds ethics.  It is not an argument which explains how God grounds ethics.  To understand how God grounds ethics we would have to first have direct perception of the divine essence, which we don't possess.  Instead, we only know the things which proceed from the divine essence, and we have to learn about what God is like, as best we can, from that.

If you like, you can take "a concrete reality which grounds ethics" as a defining property of God, and then ask questions like i) what other properties would such a being need to have, and ii) is there good reason to believe that such a being exists?

If you will allow me to make a more meta-level argument.  It seems to me that giving the Euthyphro dilemma as an objection to Classical Theism is historically obtuse.  It's like proposing the Equivalence Principle as an objection to General Relativity, when the Equivalence Principle was in fact the motivating thought experiment that led to GR in the first place.  In the same way, the question of what the gods (or really God) has to be like in order to justify treating piety as a virtue, was the underlying question motivating the Euthyphro dilemma.  But somehow atheists never say to themselves, "Geez, the fact that this famous philosophical argument was introduced in a Platonic dialogue, by a theist whose ideas laid the groundwork for the most mainstream philosophical formulation of Monotheism, maybe is a reason to think I've missed something and the argument isn't actually a knock-down in favor of Atheism."

(To be sure, arguments aren't "owned" by philosophers and there is no reason in principle why an argument by a philosopher P can't sometimes be turned against P's own worldview.  So sure, maybe there is some very subtle reason why GR is still inconsistent with the best formulation of the Equivalence Principle.  But if somebody sends me and email about why they think GR is inconsistent with the EP, and it shows no awareness of why some people have historically thought that GR satisfies the EP, then it's unlikely that their "gotcha" question about how the EP refutes GR has much merit.  Ditto for Classical Theism and Euthyprho.)

F. Metaphysical vs Logical Necessity

Now to be fair, you did explain why you don't believe in scenario C.  In addition to your "category error" assertion, you add this:

In fact, I have doubts that there is even a "metaphysical necessity" distinct from logical necessity at all. I find Chalmers' arguments in his paper "Does Conceivability Entail Possibility?" fairly convincing in this regard.

So on your recommendation, I read through this Chalmers article and I found it pretty unconvincing.  Why should reality be fundamentally scrutable to us?  Or said another way, why can't there be propositions P which are necessary, but only a mind fundamentally more powerful than the human mind could see why they are necessary?  It seems hubristic to think that human reasoning has access to every possible necessary truth.

Ironically, the reason I don't believe in Chalmers' thesis here, is actually very similar to the reasons why I side with Chalmers over Dennett when it comes to Consciousness.  While Dennett makes an interesting philosophical case for the reducibility of conscious experience to neurological facts, ultimately I concluded that Dennettism can only work if Dennettism is true by logical necessity.  In other words, that once you've specified all the physical facts then Dennett's views on consciousness follow automatically.  And it seems to me that this is simply not the case.

Similarly, Chalmers' idea that if we specify all the physical nonmodal facts, then a single set of views about modal necessity must logically follow (to idealized human reasoners) seems plainly false to me.

(Assuming it even makes sense to distinguish between "modal" and "nonmodal" facts in this way.  This is an important distinction between analytic philosophy and traditional medieval philosophy.  Analytic philosophy sees modality as primarily a feature of certain propositions, and only secondarily as a property of things.  While Aristotelian/scholastic philosophy sees modality as primarily as a property of things, while only secondarily as an attribute of propositions.  A scholastic might argue that the analytic habit of immediately jump to always reasoning about maximal "possible worlds" obscures the role that modal concepts play in causal reasoning, which involves specific concrete entities.)

Anyway, since you hold to something like Chalmers' view, here's a dilemma for you: Is the proposition expressing this view itself a logically necessary truth?

(P) There are no metaphysically necessary truths, other than logically necessary truths.

If you say that P is logically necessary, then there must be a proof that it is true which follows deductively from the definitions of the words.  What is that proof?  As far as I can tell, none exists.  Certainly Chalmers doesn't give a logically conclusive proof in that article, he just gives some reasons why he considers belief in P to be plausible, which is not the same thing.

On the other hand, if is not logically necessary, then either it is contingent (which is inconsistent with the usual S5 rules for modal logic) or else it is an example of a metaphysically necessary (but not logically necessary) truth, in which case it refutes itself.

One could make a similar, superficially less "meta" argument for the same conclusion by considering the proposition:

(N) A necessary being exists.

A standard analytic argument from S5 modal logic implies that either: i) N is necessarily true, or ii) N is necessarily false.  So which of these is logically necessary?  I say neither, but if you disagree then what do you think the proof of N or its negation would look like?

G. Can God be the grounds of Logic?

I believe that you've stated elsewhere that while you believe God is metaphysically necessary, he is not logically necessary - but of course, it is logically necessary that the laws of logic or mathematics are true. I don't think the dependence you're arguing for could work, even if God exists in some sense.

This is a little compact, but I'm guessing your argument is something like the following:

1. A contingent truth cannot ground a necessary truth.*
2. God's existence is logically contingent.
3. But logic itself is logically necessary,
4. Therefore, God cannot ground logic.

[*I suppose there is some sense in which, if a Cat walks onto a Mat, this arguably grounds the necessary proposition: "Either the Cat is on the Mat or the Cat is Not on the Mat" by virtue of being a truthmaker for one of its disjunctives.  But I won't pursue this possible counterexample further, since I don't think it is relevant to the sense in which God grounds logic.]

But this argument is fallacious, because when I say that God grounds logic, I am making a metaphysical statement rather than a logical one.  From the perspective of metaphysics, both logic and God are (in my view) metaphysically necessary, and it is not at all impossible for a necessary statement to ground another necessary statement.  In other words, we have to distinguish between:

1a: A logically contingent truth cannot logically ground a logically necessary truth.

which is true, and:

1b: A logically contingent truth cannot metaphysically ground a logically necessary truth.

which does not in any way follow from 1a, and I would say it is false.

H. What Metaphysical Necessity Means

Actually, there is a better way to put this which makes the concept of "metaphysical necessity" somewhat less mysterious.  The right way to talk about this is to make Aristotle's distinction between that which is necessary to us (axioms of human thought) and that which is necessary in itself (propositions which could not have been otherwise).

When we say that a proposition is metaphysically necessary, we merely mean it falls into the latter category.  The adjective is misleading since, unlike the cases of "logical necessity" or "nomic necessity" (which mean necessary given certain specific principles), the phrase "metaphysically necessary" simply means whatever is necessary simpliciter, i.e. that which (without adding any qualifications) could not have been otherwise (whether or not the reason for its necessity is known to human beings.)

On the other hand, logical necessity is an example of what is necessary to human beings, i.e. an axiom of human reasoning, or a particular technique L used to prove the impossibility of certain propositions.

So, the proposition P from earlier boils down to:

(Equivalent to P): If a proposition cannot be proven to be impossible by technique L, then it really is possible.

while I see no reason to believe that technique L is sufficient to uncover all possible cases of necessity.  Especially since technique L does not even seem to be powerful enough to refute the statement that no concrete entity whatsoever exists.

This relates of course to cosmological considerations as well.  As is well-known, if P is true, then the basic principles of existence are just contingent "brute facts" which means they are not true for any reason at all.  So there is an obvious reason to postulate a necessary concrete entity, which is that it serves as a starting point to explain why anything else exists at all.

This reason to want a necessary being, does not seem to depend on us being able to know why the being is necessary.  This is the Thomistic viewpoint on the Cosmological Argument, and it seems to me to be the only possible middle ground between Anslemian positions (there is a valid Ontological Argument for a necessary being from pure logic) and explanatory nihilism (there is no good reason why the universe exists, it just does).

(Now you could just double down and say, I have no idea what you mean by the phrase: ``could not be otherwise'', please explain it to me; and then refuse to accept any answer I give other than one which reduces it to logical implication.  But the same technique could be done to motivate skepticism towards practically any other concept, including the other concepts in this discussion like "mind" or "good" or "abstract" or "grounds".   (It is not even clear that logical necessity can be fully explained without an infinite regress, as  St. Lewis Carroll pointed out in his Achilles and the Tortoise dialogue.)  I don't claim to have a definition of metaphysical necessity that would satisfy Socrates, but if we make that the standard, there aren't going to be very many philosophical terms left!)

I. An Irrelevant Topic

As a final statement, I don't think theism is actually better at convincing people of being moral than secularism.

This is just so totally irrelevant to the metaphysical questions behind the Moral Argument for Theism, that perhaps I should simply refuse to respond to this entirely.  It's really just a complete change of topic.

God could be the metaphysical grounds for morality, even if every single human being on Earth were an atheist, or even if every single theist were morally worse than every single atheist.  These motivational questions really have nothing whatsoever to do with the question about what metaphysical theses are made more plausible, if we subscribe to moral realism.  I wrote my blog post Is it Possible to be Good without God? precisely because I was annoyed by how regularly people seem to conflate these totally unrelated questions.

(I'm not saying that the degree of goodness of religious people can't potentially be used as an evidential argument for or against the existence of God.  What I am saying is that it is a mistake to allow such sociological questions to contaminate our interpretation of the thesis that God grounds ethics.)

That being said, I''ll take the bait and say I do think there is some pretty serious question begging required for a non-circular argument that atheism is fully compatible with moral behavior.  For one thing, if a being such as is described by Classical Theism in fact exists (a perfectly wise and holy and good being, who created us and is the source of all our goodness), then we have the moral obligation to worship and obey that being, and to reflect God's holiness through a life of prayer and repentance, dedicating our earthly activities to the glory of God.  It is difficult to see how an atheist can satisfy that obligation, because for the atheist these activities are just distractions from a different, more secular understanding of what the good life consists of.

(To be sure, if the atheist has some intellectually honest reasons why they think God does not exist, then this may well be a mitigating circumstance that reduces—or even eliminates entirely—their culpability for this omission.  But if we are discussing the question of which beliefs make it easier to be moral, then usually mitigating circumstances are considered mitigating precisely because they make it harder to be moral.  Furthermore, a lack of culpability does not remove all of the causal consequences of trying to place our ultimate happiness in things other than God—what Christians call idolatry.)

I do suspect nonreligious people being more moral than the religious, if true, would be a particularly big problem for theism and theistic morality.

From the standpoint of Christian doctrine, it is not actually clear why this should be.  Merely having knowledge of God's existence does not necessarily translate into obedience, and in some cases knowledge can make people morally worse since they ought to behave better but don't.  As Jesus' brother St. James said:

You believe that there is one God. You do well. Even the demons believe—and tremble!  (James 2:19)

and as Jesus himself said:

“Not everyone who says to me, ‘Lord, Lord!’ will enter the kingdom of heaven, but only the one who does the will of my Father in heaven.  On that day many will say to me, ‘Lord, Lord, didn’t we prophesy in Your name, drive out demons in Your name, and do many miracles in your name?’  Then I will announce to them, ‘I never knew you! Depart from me, you lawbreakers!’ ”  (Matthew 7:21-23)

The Pharisees were among the most "religious" people in Jesus' day, and many of their leaders handed Jesus over to Pilate to be crucified.  See also St. Paul's observations of religious people in Romans 2.

According to Christianity, what people need to be transformed morally, is not so much knowledge as grace.  Knowledge is good if it helps us acknowledge our need for grace, but not so much if it makes us look down on other people.

I think the evidence at least shows that the nonreligious are generally not less moral than the religious...

I'm not sure what evidence you are referring to here, or how you could actually know this to be the case.  If your claim is just that religious people can be morally weak and inadequate, well I already knew that from my own life, without looking at anybody else's.

If it refers to survey data, you have the problem that what many polls of religious affiliation captures a lot of individuals who only identify as religious in a nominal sense.  Polling nominally religious people, and asking about their rates of divorce, adultery etc. is sort of like asking whether watching the Olympics on TV makes people more physically fit!  It's the wrong question to study.

If you are referring to personal experience, I can only say that while I know good and bad seeming people (emphasis on "seeming", it's not my place to judge them) who are both religious and non-religious, the most loving and self-sacrificial people I know seem to be religious.  And religion also often plays a significant role when very bad seeming people repent and turn their lives around.  Furthermore I have very often heard people refer explicitly to God when they explain why they did something morally difficult, while I cannot ever recall in my personal experience ever hearing somebody say that they did something morally difficult because atheism is true.  (I mean, I could imagine such a motivation: e.g. God isn't going to save this person, so I have to.  But I don't think I've ever heard anyone explicitly say this "in the wild" so to speak.)

By comparison, studying secular ethics seems to itself have little observable consequences in terms of making people better.  This could be taken as a critique of secular ethics, but it might be better taken as a critique specifically of what modern analytic philosophers mean by ethics as a discipline (as opposed to ancient philosophies, which were typically viewed as a way of life that had to be put into practice, in order to be understood).  I mean, why should studying little numbered arguments about whether ethics is objective, or arguing about what to do in some controversial edge case involving trolleys, actually help one to build habits of life that make one treat your fellow human beings better, and a community which helps support you in doing so?  Religion is one of the few ways of getting such support in the modern era.  (There are some others, but they are getting sparser in an increasingly disconnected age.)  While this isn't necessarily an argument for God's existence, it does make your thesis that serious religious practice is totally orthogonal to ethical accomplishment seem pretty implausible.

I called this an "irrelevant topic" because it isn't terribly relevant to the validity of the Moral Argument.  But of course, from the perspective of what ultimately matters, it is this section that is most important, and the rest which are of lesser relevance.  If Christianity is true, then what will matter the most in the end is not whether you are persuaded by this or that specific argument for Theism, but more whether your heart is open or closed to God at a deeper level than that.  Jesus has promised that those who truly seek God will find him.

If you take it as a goal to be as moral of a person as you can possibly be, then that is at least a start along that road—even if the final destination is going to be, in some ways, quite different than what you expected when starting out on that journey.  But somewhere along the way comes the recognition that you can't actually be good, and need help to do better, and that is where concepts like grace and salvation start to make more sense...

Blessings,
Aron

Posted in Ethics, Metaphysics, Theological Method | 40 Comments

In the Valley of the Shadow of Death

A lesson from Martin Luther on walking in faith during pandemics.

By Scott Church – Guest Blogger

After decimating nearly one-third of Europe during the 14th Century, the Bubonic plague continued to ravage it in periodic epidemics before it was effectively eradicated in the mid-20th Century (White, 2014; Schiferl, 1983; Griggs, 2014). For the most part, these outbreaks were isolated to villages or regions, and it was possible to flee to safety elsewhere until they subsided. In August of 1527, one such outbreak came to Wittenberg while Martin Luther was at the university there, and Elector Johann Hess of Saxony ordered him and other professors to flee to Jena for safety.

Luther refused, choosing instead to stay behind with his wife Katharina von Bora and open their home as a ward for the sick, whom they cared for at great personal risk to themselves. He penned a letter to Elector Johann explaining his reasons (Luther, 2020). Five centuries later, in the age of COVID-19, his words and the testimony of his life show us what true God-fearing faith during pandemics is... and more importantly, what it is NOT.

In his words,

"[W]hoever serves the sick for the sake of God's gracious promise... has the great assurance that he shall in turn be cared for. God himself shall be his attendant and his physician, too. What an attendant he is! What a physician! Friend, what are all the physicians, apothecaries, and attendants in comparison to God? Should that not encourage one to go and serve a sick person, even though he might have as many contagious boils on him as hairs on his body, and though he might be bent double carrying a hundred plague-ridden bodies! ... Therefore, dear friends, let us not become so desperate as to desert our own whom we are duty-bound to help and flee in such a cowardly way from the terror of the devil, or allow him the joy of mocking us and vexing and distressing God and all his angels..."

True disciples don't deliberately put themselves in harm's way out of mere fealty to church doctrine, or to appease worldly narratives and political agendas others have tarnished it with for reasons that serve their own interests rather than God's. They do so in loving service to their neighbor. In the words of the apostle Paul, they offer themselves as living sacrifices, holy, acceptable to God, which is their reasonable service (Rom. 12:1).

Note the reference to reasonable service (from the KJV Bible)—Or as the Amplified Bible renders it, "rational (logical, intelligent) act of worship." Genuine faith sees the face of Jesus in the poor, the oppressed, and the sick, and with full rational knowledge of the risks involved, seeks to be His healing face in their lives. It is in THAT place that we trust God's Will for our best, and our neighbor's.

By contrast, Luther tells us, there are those who,

"Sin on the right hand. They are much too rash and reckless, tempting God and disregarding everything which might counteract death and the plague. They disdain the use of medicines; they do not avoid places and persons infected by the plague, but lightheartedly make sport of it and wish to prove how independent they are. They say that it is God's punishment; if he wants to protect them he can do so without medicines or our carefulness."

This sort of "faith" will have nothing to do with reason (logic, intelligence). It flies recklessly in the face of real-world facts, rejects medicine, makes no attempt to socially distance from the sick, and even goes so far as to make fun of those who do, simply to assert its independence... that is, freedom.

Sound familiar...?  ;-)

According to Luther,

“This is not trusting God but tempting him. God has created medicines and provided us with intelligence to guard and take good care of the body so that we can live in good health… If one makes no use of intelligence or medicine when he could do so without detriment to his neighbor, such a person injures his body and must beware lest he become a suicide in God's eyes. By the same reasoning a person might forego eating and drinking, clothing and shelter, and boldly proclaim his faith that if God wanted to preserve him from starvation and cold, he could do so without food and clothing. Actually that would be suicide.

It is even more shameful for a person to pay no heed to his own body and to fail to protect it against the plague the best he is able, and then to infect and poison others who might have remained alive if he had taken care of his body as he should have. He is thus responsible before God for his neighbor's death and is a murderer many times over. Indeed, such people behave as though a house were burning in the city and nobody were trying to put the fire out. Instead they give leeway to the flames so that the whole city is consumed, saying that if God so willed, he could save the city without water to quench the fire..."

True disciples are rational (logical, intelligent). They embrace science, medicine, and socially responsible behavior—not out of license masquerading as "freedom," but because they are responsible to God for their own health, and... [wait for it] ... their neighbor's. To do otherwise—to reject their reasonable service, which is holy, acceptable to God—is to tempt Him rather than trust Him, and in so doing, become a murderer plain and simple.

In summary, he tells us,

"No, my dear friends, that is no good. Use medicine; take potions which can help you; fumigate house, yard, and street; shun persons and places wherever your neighbor does not need your presence or has recovered, and act like a man who wants to help put out the burning city. What else is the epidemic but a fire which instead of consuming wood and straw devours life and body? You ought to think this way: Very well, by God's decree the enemy has sent us poison and deadly offal. Therefore I shall ask God mercifully to protect us. Then I shall fumigate, help purify the air, administer medicine, and take it. I shall avoid places and persons where my presence is not needed in order not to become contaminated and thus perchance infect and pollute others, and so cause their death as a result of my negligence. If God should wish to take me, he will surely find me and I have done what he has expected of me and so I am not responsible for either my own death or the death of others. If my neighbor needs me, however, I shall not avoid place or person but will go freely, as stated above. See, this is such a God-fearing faith because it is neither brash nor foolhardy and does not tempt God."

Nor is this restricted to personal faith only. It is a calling to the church and community as well,

“'Whoever loves danger,' says the wise man, 'will perish by it' (Ecclus. 3:26). If the people in a city were to show themselves bold in their faith when a neighbor's need so demands, and cautious when no emergency exists, and if everyone would help ward off contagion as best he can, then the death toll would indeed be moderate. But if some are too panicky and desert their neighbors in their plight, and if some are so foolish as not to take precautions but aggravate the contagion, then the devil has a heyday and many will die. On both counts this is a grievous offense to God and to man..."

To these ends, Luther’s exhortation to “make use of medicine and intelligence” is particularly timely for us. When diseases broke out in his world, one had only two options—do your best to avoid them; and pray for a healthy recovery if you don’t succeed. We, on the other hand, have been blessed with five centuries of advances in virology, immunology, and medicine his world didn’t have. And of all the blessings at our fingertips in the age of COVID-19, one stands out more than any other—the one that allows us to arm ourselves against it, and possibly even eradicate it… vaccines. Unfortunately, many people still aren’t getting them, which is keeping widespread herd immunity out of reach. In the United States in particular, many are flat-out refusing vaccination for ideological reasons, not the least of which is a general hostility toward science and public health measures that from all appearances, no amount of evidence or logic will ever be able to penetrate. Many others, however, are hesitant due to concerns about how safe and effective COVID vaccines are (especially considering public health recommendations to continue masking and social distancing even after vaccination) but can otherwise be reasoned with if these concerns are addressed. They can be.

COVID vaccines are effective

As of this writing, three COVID-19 vaccines are in general use in the United States: The messenger RNA-based (mRNA) vaccines manufactured by Pfizer and Moderna, and the Johnson & Johnson adenovirus-based "one-shot" vaccine. All three have been thoroughly tested and approved by the FDA (Tanne, 2020; Oliver, 2020). The AstraZeneca adenovirus-based vaccine has also been approved for general use in Europe (EMA, 2021). Demonstrated efficacies of mRNA-based vaccines against infection or symptoms requiring hospitalization from the original wild strains of SARS-COV-2 are 95-97% for the Pfizer–BioNTech BNT162b2, and 92-95% for Moderna mRNA-1273. Corresponding figures for the Johnson & Johnson [J&J] Ad26.COV2.S and AstraZeneca–Oxford ChAdOx1 nCov-19 vaccines are around 67-72% (Haas et. al., 2021; Tenforde et. al., 2021; Callaway, 2021; Noor, 2021; Polack et. al., 2020; Mahase, 2020; Olliaro et. al., 2021; Mallapaty & Callaway, 2021).

As of Sept. 2021, these figures are still holding up well, even against recent variants such as B.1.617.2, or Delta. Per multiple studies in Europe and North America, effectiveness of the Pfizer–BioNTech vaccine against the more robust and transmissible Delta variant ranges from 79% to 88% for infection and symptomatic illness, and 89% to 100% (!) for hospitalization (Tregoning et. al., 2021; Lopez Bernal et. al., 2021; Baraniuk, 2021; CDC, 2021).

For all vaccines collectively, one recent study in New York found overall age-adjusted effectiveness against new COVID-19 cases and hospitalizations to be 75% and 89.5% to 95.1% respectively (Rosenberg et. al., 2021). A similar recent study in England found 50-60% effectiveness against infection by Delta (symptomatic or otherwise), including the less effective one-shot ones such as J&J (Smout, 2021). Even a single immunization has been shown to boost neutralizing titers against all variants and SARS-CoV-1 by up to 1000-fold (Stamatatos et. al., 2021), and one study of new COVID-19 cases in Kentucky during May and June of 2021 found that those who were vaccinated were 2.34 times less likely to be infected than those who had previously had COVID-19 and survived but weren't vaccinated (Cavanaugh et. al., 2021). One recent study in Israel did find an effectiveness of only 64% for Pfizer–BioNTech BNT162b2 against infection and symptomatic illness (Hass et. al., 2021). However, it was based on incidence rates in subjects who were considered fully vaccinated one week after receiving their second dose, whereas per U.S. CDC guidelines, one isn't considered fully vaccinated until two weeks after their second dose (CDC, 2021b).

If one does contract COVID-19 after vaccination, severe symptoms, hospitalizations, and deaths among breakout cases are almost an order of magnitude lower than those among the unvaccinated. Even in the case of the more vaccine-resistant Delta variant, the Pfizer–BioNTech BNT162b2 and Moderna mRNA-1273 vaccines reduce risk of hospitalization after four months by 93% and 91% respectively, and by 92% and 77% after six months (Scobie et. al., 2021; Self et. al., 2021).

But of course, if in doubt one could simply check the trended data on new US cases and deaths vs. vaccination rates since mass distribution of these vaccines began in earnest last January (JHUM, 2021). The dramatic declines in COVID-19 with rising national vaccination levels reflected in these datasets are self-evident. The spike in new cases after July 11, 2021 was almost entirely due to the Delta variant spreading among the unvaccinated, who as of July 30, 2021 comprised 96-99.8% of all cases (Kates et. al., 2021). And among the rising percentages of breakthrough cases (thanks to the unvaccinated Petrie dish), severe illness, hospitalizations and deaths are clearly a fraction of those for the unvaccinated (CDC, 2021c; Evans & Wernau, 2021).

By the numbers and the extensiveness with which they've been tested, the effectiveness of these vaccines in preventing infection, hospitalization, or death from COVID-19 is beyond reasonable dispute. But that said, it's important to be clear about what we mean by effectiveness and efficacy (there's a difference). When we say, for instance, that the Pfizer–BioNTech vaccine has an efficacy of 88% against infection, we mean that in controlled studies where a random sample of subjects received the Pfizer vaccine and an identical (or as similar as possible) control group of subjects received a placebo, 88% fewer subjects in the vaccinated group contracted COVID-19 during the trial period than the unvaccinated group--that is, if 100 COVID-19 cases turned up in the unvaccinated group, twelve did in the vaccinated group, and likewise for efficacies against hospitalization and death. On the other hand, vaccine effectiveness generalizes these comparisons to wider vaccine use in the general public. Since vaccine distribution and use may differ regionally and/or demographically from controlled laboratory studies, vaccine effectiveness may differ somewhat from efficacy.

In both cases, what we are NOT saying is that an efficacy/effectiveness of 88% against infection means that vaccines only work for 88 out of 100 people, nor that they will only work 88% of the time for you. Likewise, 93% efficacy/effectiveness against hospitalization does NOT mean that seven out of every 100 breakout cases will be hospitalized, and the rest will be asymptomatic. It isn't hand grenades. :-)

It simply means that there will be 88% fewer infections and 93% fewer hospitalizations in a vaccinated population than an unvaccinated one. But everyone who is vaccinated still has some level of protection from vaccines that they wouldn’t otherwise have. [The WHO Vaccine efficacy, effectiveness and protection page has a very readable and informative overview of all this.]

All other factors held constant the bottom line is that vaccination protects everyone and does so in at least three ways.

First, while it is true that in some cases the individual protection offered by vaccines may not be enough to prevent one from coming down with the disease or being hospitalized, they still reduce everyone’s risk for infection, and nearly all of those who do come down with a breakout case anyway will have less severe symptoms than they otherwise would have. How well vaccination protects you personally will depend on a wide range of factors including your age, your overall immune function, any comorbidities you may have, how much exposure you get from daily life (home, workplace, etc.), and more. But regardless, you will be more protected with vaccination than without it. And unless you have known life-threatening vaccine allergies or related immune function risks, getting vaccinated poses no risk compared to remaining unvaccinated since you would have to be infected and get sick to generate an immune response anyway, so there's no reason not to get one.

Second, if 88 out of 100 people who are vaccinated don’t contract COVID-19 when exposed to it, that means there are 88 fewer people spreading the disease before they develop symptoms, which in turn reduces everyone’s risk of exposure to it in the first place (more on this shortly). This is a key point, especially for those who intend to love their neighbor as themselves…

Choosing to be vaccinated doesn’t just protect you from infection, it protects your loved ones, your friends, and your community.

Finally, and most alarmingly, the vast majority of people filling hospital beds nationwide and around the world are unvaccinated COVID-19 patients, and the resulting burden is taxing healthcare workers and resources to the breaking point—so much so that in many regions, hospitals are literally having to resort to “death panels” to decide who gets care based on their likelihood of survival (Knowles, 2021; Hiltzik, 2021; Westneat, 2021). In other words, we have now reached a point in this pandemic where people are literally dying from preventable conditions because there are no hospital beds for them.

A month ago, my 89-yr-old father fell and broke his knee. He was left on a gurney in a hallway at Deaconess Hospital in Spokane, Washington for eight hours because there wasn’t a single bed available for him—all but a handful were being used by unvaccinated COVID-19 patients from Idaho who were seeking care in Washington because of the very Idaho hospital death panels discussed in the last two sources cited above. If he’d been in a car accident, needed an emergency appendectomy, or had a heart attack, he’d be dead… for literally no reason other than that all the beds in the nearest hospital were taken up by unvaccinated COVID-19 patients.

Choosing to be vaccinated doesn’t just protect you from hospitalization and death, it protects doctors, nurses, and healthcare workers struggling to save lives, and saves everyone from needless crippling or death due to lack of available care.

COVID vaccines are safe

As of this writing, nearly 6.3 billion COVID-19 vaccinations have been administered worldwide. More than 393 million have been administered in the United States, and 63% of the U.S. population have had at least one shot (Ritchie et. al., 2021; JHUM, 2021). Anaphylaxis adverse reaction rates have run around 0.0011% for Pfizer and 0.00025% for Moderna or roughly two to eleven adverse events per million vaccinations administered (Rutkowski et. al., 2021; Shimabukuro et. al., 2021; Banerji et. al., 2021). Corresponding figures for adenovirus vaccines such as Johnson & Johnson [J&J] Ad26.COV2.S and AstraZeneca–Oxford ChAdOx1 are around 0.0003% for blood clotting (Ledford, 2021; CDC, 2021d). Overall, as of Aug. 16, 2021, after administration of more than 357 million doses of COVID-19 vaccines, a grand total of 6,789 deaths had been reported, or 0.0019% of doses administered (CDC, 2021d), and few of these deaths have even been specifically tied to the vaccines themselves rather than extraneous factors or even coincidence. For these and many other reasons, as of Aug. 23, 2021, the Pfizer–BioNTech BNT162b2 has full rather than emergency FDA approval (USFDA, 2021).

For comparison, your odds of being struck by lightning once in an 80-year lifetime (believe it or not, the National Weather Service maintains stats on this!) are one in 15,300, or 0.0065%--more than three times the odds of a severe adverse reaction (SAR) from any COVID-19 vaccine (NWS, 2021). Apart from valid doctor-certified medical exemptions, it isn’t reasonable to refuse vaccination based on risk this low.

In conclusion, it should be also noted that there is a flood of disinformation regarding vaccine safety and effectiveness circulating on social media and in online activist and news/op-ed forums. A detailed examination of the numerous claims and allegations being made is beyond our scope today but suffice to say that virtually none of it has any basis whatsoever in fact and it continues to spread only because it receives uncritical reception in these forums outside of the scientific peer-review process.1 By the reliable data and numbers, the safety of these vaccines is also beyond reasonable dispute.

Do I still need to mask up and socially distance after vaccination?

In a word, yes… but only as the circumstances of your daily activities and regional safety guidelines dictate. Here are the things that need to be kept in mind…

As of this writing, 99% of all new COVID-19 cases in the US are the Delta variant (CDC, 2021e). As already noted, the existing Pfizer vaccine has been shown to be 79-88% effective against Delta for infection. That's tantamount to saying that it's 12-21% ineffective, meaning that even if you're vaccinated you still have roughly one chance in six of coming down with COVID-19 if exposed to it, perhaps asymptomatically.

What happens if you do...? It’s well known that breakout cases among the vaccinated can still carry viral load significant enough to be contagious even if they don't become symptomatic, and in some cases, they may even carry as much as those who aren’t vaccinated (CDC, 2021). Either way, if you do, how many susceptible people you pass it to while contagious will depend on a wide range of factors—your age and immune function, demographics of your daily encounters, behavior (including masking and social distancing), etc. Taking all these factors into account, given the average person infected with a disease, the expectation value for how many people he/she will spread it to in an unvaccinated environment while contagious is given by its base reproductive factor, or R_{0}.

As of this writing, Delta has an estimated R_{0} of between 5 and 9.5, as opposed to that of chickenpox, which has an R_{0} of 8.5 (CDC, 2021; UNSW, 2021; Liu & Rocklöv, 2021; Georgiou, 2021). As such, even if you are vaccinated, if you come down with a breakout case of Delta COVID-19 in an unvaccinated setting and don't quarantine or change your behavior, you will likely spread the disease to at least some people before recovering or dying. In most cases being vaccinated will reduce the likelihood that you will spread it, but it’s possible that you could spread it to as many as five to nine others. Each of them will then do likewise, and so on—more so among the unvaccinated. As successive generations of infection proceed through a given population, the number of susceptible hosts will be eroded by acquired immunity or death, and continued infection rates will to first order yield an effective reproductive factor, R_{eff}, given by,

R_{eff} = R_{0}\left ( 1 - p_{1} \right )

where p_{1} is the percentage of a population that has acquired immunity either through infection or... vaccination. As can be seen, the key to reducing R_{eff} is to increase p_{1}… And vaccination makes this possible at a much faster rate with orders of magnitude fewer casualties.

For Delta (or any other SARS-COV-2 variant) to be contained regionally or globally, R_{eff} must remain less than 1.0 long enough for the virus to die out. So, given a median R_{0} of 7.3 for the estimated range above, this means that p_{1} must be greater than 0.86. As of Oct. 3, 2021, total cumulative U.S. COVID-19 cases were at 43.7 million and deaths at 701,000, or around 13.1% of its population that has acquired immunity, and concurrently, 54.9% of its population is fully vaccinated (JHUM, 2021; CDC, 2021). Conservatively assuming negligible breakout case overlap, and naively presuming a normalized overall vaccine effectiveness of 88% (per the upper range of Pfizer–BioNTech Delta variant effectiveness cited above), that works out to at most, a p_{1} of 0.61—far short of the target needed for containment. And none of this accounts for the erosion of vaccine effectiveness by the evolution of increasingly vaccine-resistant strains, which once they break out of vaccinated hosts, spread most virulently among the unvaccinated.

What can we do? By my lights, there are three responsible options:

Option #1:  If you haven’t done so already, consider getting vaccinated.

This is by far, the best protection you can offer yourself and others against infection and/or hospitalization from all extant strains of SARS-COV-2. If you have a history of allergies and/or reactions to vaccines and are worried about whether they’re safe for you, consult your primary care doctor. You might also want to spend some time at the CDC’s COVID-19 Vaccine Information portal for more information. Bear in mind that these vaccines are free. You don't need health insurance to get them and they’re available at most pharmacies as well as clinics, including grocery store pharmacies (My wife and I got both our Pfizer shots at our neighborhood Safeway). The pharmacists there will gather the needed information regarding your risks, and consult your primary care doctor as well if need be. For safety reasons, you will be asked to remain in the waiting area for 10-30 minutes after receiving your shot. And in the extremely unlikely case that you do have a SAR (Severe Adverse Reaction) to vaccination, they will have EpiPen’s on hand that will immediately rectify all but the tiniest handful of them.

Again, this cannot be emphasized enough—There is an obscene amount of pseudoscience, conspiracy theories, and other disinformation being circulated on social media by anti-vax activists. 1 To repeat a viral mantra in these communities… Under no circumstances whatsoever should you “do your own research” on YouTube, Facebook, or any agenda-driven online forums outside of the scientific peer-review process. Your primary care doctor has your personal medical history, and properly trained pharmacists who work with COVID-19 vaccines and understand what risks they have will be able to contact him/her if there are any concerns. They and they alone can speak to whether they’re safe for you.

Option #2:  Mask and socially distance when prudent, especially indoors.

If COVID-19 vaccines aren’t a safe and viable option for you, you can still protect yourself and others by socially distancing and wearing a mask. SARS-COV-2 is spread primarily by expectorated droplets and aerosols (this is where the six-foot rule comes from) and masking dramatically decreases the spread of these droplets. Outdoors, breezes and atmospheric dispersion make this less of a concern. But indoors it’s more important, especially in smaller spaces.

The best protection is provided by medical-grade N95 masks like those manufactured by 3M’s Particulate Respirator 8211. These are the only masks that will individually block SARS-COV-2 viral transmission in both directions, protecting you as well as others. Their only downsides are limited availability, and for some people, discomfort (they tend to produce skin irritation and/or itching).

The next best thing is a high-quality 3-ply cloth mask with microfilters such as those made by Airband. Even better is double-masking—wearing a surgical mask under a 3-ply cloth one. Recent research has shown that properly done, this can reduce one’s risk of transmission and infection by 90% or more, rivaling the efficacy of mRNA vaccines (Brooks et. al., 2021). Proper use of masks is as important as mask selection, so it’s a good idea to review the CDC’s Guidelines for improving mask protection.

It also should be pointed out that agenda-driven activists on social media and in online “news” and propaganda forums are spreading even more pseudoscience and disinformation about masks and social distancing than vaccines, and virtually none of it has any basis whatsoever in fact either. 2 As before under no circumstances whatsoever should anyone be “doing their own research” in such forums outside of the scientific peer-review process.

Option #3:  Avoid crowds and prolonged indoor gatherings.

As already noted, expectorated droplets are the primary vector of transmission for SARS-COV-2. However, normal breathing does release a viral load that only a medical-grade N95 mask will stop. In outdoor or large, well-ventilated spaces this viral load is too small to make a difference. But in tightly crowded conditions and gathering in small, enclosed spaces it can build up to dangerous levels. If you don’t have access to medical-grade N95 masks, avoid crowded gatherings in poorly ventilated spaces—yes, unfortunately, that does include churches where proper circulation and social distancing measures aren’t being implemented.

Finally, bear in mind that as we have seen, even if you are vaccinated, adopting options #2 and #3 as well will still give you protection from a breakout infection, and help protect others if you do come down with one.

Whatever path we choose, let us examine our own hearts and remember that it’s not just we ourselves that we’re protecting, but our neighbors, our loved ones, and our communities. As the poet John Donne said,

“No man is an island entire of itself; every man is a piece of the continent, a part of the main; if a clod be washed away by the sea, Europe is the less, as well as if a promontory were, as well as any manner of thy friends or of thine own were; any man's death diminishes me, because I am involved in mankind. And therefore never send to know for whom the bell tolls; it tolls for thee.”

As we face our own plague—As millions of our fellow citizens suffer under the iron fist of this cruel disease, hundreds of thousands die slow, horrible, intubated deaths, and doctors and nurses put in 70/80-hour weeks at the edge of their human reserves to save lives—Luther reminds us that we are all in this together, and we’ve been called to go forth into that Valley of the Shadow of Death hand-in-hand...

Not in brashness or foolhardiness… Not in willful rejection of science and medicine… Not in service to Self and license masquerading as "freedom…"

But as living sacrifices, holy, acceptable to God, in reasonable service to each other knowing that whatever may befall us, God is by our side completing the work he began in us. "Truly I tell you, whatever you did for one of the least of these brothers and sisters of mine, you did for me" (Matt. 25:39-40).

Or in the words of Paul,

"All things are lawful, but not all things are profitable. All things are lawful, but not all things edify. Let no one seek his own good, but that of his neighbor" (I Cor. 10:23-24).

"Do nothing from selfishness or empty conceit, but with humility of mind regard one another as more important than yourselves; do not merely look out for your own personal interests, but also for the interests of others. Have this attitude in yourselves which was also in Christ Jesus, who, although He existed in the form of God, did not regard equality with God a thing to be grasped, but emptied Himself, taking the form of a bond-servant, and being made in the likeness of men. Being found in appearance as a man, He humbled Himself by becoming obedient to the point of death, even death on a cross. For this reason, also, God highly exalted Him and bestowed on Him the name which is above every name, so that at the name of Jesus every knee will bow, of those who are in heaven and on earth and under the earth, and that every tongue will confess that Jesus Christ is Lord to the glory of God the Father" (Phil. 2:3-11).

Do nothing from selfishness or conceit… Regard others as more valuable than yourself, and look to their interests as well as your own...

Have this attitude (this mindset, this worldview, these values... not these parroted dog-whistles or party-line narratives) in you which was also in Christ Jesus...

Who although He was God Incarnate, with all the power, authority, and glory thereof, did not consider that august status a thing to be grasped (clung to, defended with bared teeth and narcissistic injury), but emptied Himself, taking on the role of a servant...

And being found in mortal human form, was obedient to the point of death—even death on a cross, which in New Testament times was a death of disgrace reserved only for the lowest of despised criminals...

This is the kind of discipleship we’ve been called to… And it’s a far cry from rugged individualism and idolatrous nationalism whitewashed with joyful hymns and inspirational bumper stickers.

Say what you will about his quaint puritanical language, his belief that "evil spirits" cause plagues, and other bucolic naivetes. But like us, Luther was a man of the age he lived in. His words were penned long before he or any of his contemporaries had access to modern epidemiology, immunology, or even knowledge of germs. To dismiss him for speaking from, and to the age he lived in would be at best anachronistic, and at worst, sanctimonious. Archaic or not, in this age of COVID-19, the example he left with us is as self-evident as it is timeless, and those of us who call ourselves Christians would do well to heed it—especially those who seem to think that trusting God means tempting Him by rejecting science and medicine and behaving recklessly in the name of “freedom,” and then expecting Him to clean up their messes without holding them accountable as His sons and daughters.

We can embrace a faith like his that "makes use of intelligence and medicine" and "serves the sick for the sake of God's gracious promise." We can offer ourselves as living sacrifices, holy, acceptable to God in reasonable service to our fellow human beings and put an end to this pandemic. We can reach for the best that is in us, the best that is in our souls...

Or we can set aside loving our neighbors as ourselves (Mark 12:31) and tempt God with a "faith" based on denial, recklessness, and idolatrous worldly narratives and spread this disease throughout the world, filling hospitals and graves in our wake.

In short, we can be salt and light to a world in need... or in Luther's words, murderers.

The choice is ours. But make no mistake... We're kidding ourselves if we think we can choose the latter and expect that outside of our own echo chambers, the world isn't going to notice the difference and judge our witness accordingly.

Footnotes

1)      A deeper examination of some of the most widespread anti-vax myths currently in circulation can be found at two public Facebook posts of my own titled Covid-19 Vaccine Whack-A-Mole and Covid-19 Vaccine Whack-A-Mole - Part 2.

2)      Likewise, a deeper examination of the most widespread anti-mask myths currently in circulation can be found at a public Facebook post of my own titled Anti-Mask Whack-A-Mole.

References

Banerji, A., Wolfson, A.R., Wickner, P.G., Cogan, A.S., McMahon, A.E., Saff, R., Robinson, L.B., Phillips, E. and Blumenthal, K.G., 2021. COVID-19 Vaccination in Patients with Reported Allergic Reactions: Updated Evidence and Suggested Approach. The Journal of Allergy and Clinical Immunology in Practice. Online at https://www.jaci-inpractice.org/article/S2213-2198(21)00466-9/abstract. Accessed Oct. 3, 2021.

Baraniuk, C., 2021. Covid-19: How effective are vaccines against the delta variant? BMJ: British Medical Journal, 374. Online at https://www.bmj.com/content/374/bmj.n1960. Accessed Oct. 3, 2021.

Cavanaugh A.M., Spicer K.B., Thoroughman D., Glick C., & K. Winter. 2021. Reduced Risk of Reinfection with SARS-CoV-2 After COVID-19 Vaccination — Kentucky, May–June 2021. MMWR Morb Mortal Wkly Rep, 2021;70:1081-1083. DOI: http://dx.doi.org/10.15585/mmwr.mm7032e1. Online at https://www.cdc.gov/mmwr/volumes/70/wr/mm7032e1.htm?s_cid=mm7032e1_w. Accessed Oct. 3, 2021.

Centers for Disease Control and Prevention (CDC). 2021. Improving communications around vaccine breakthrough and vaccine effectiveness. PowerPoint presentation, July 29, 2021. Online at https://context-cdn.washingtonpost.com/notes/prod/default/documents/8a726408-07bd-46bd-a945-3af0ae2f3c37/note/57c98604-3b54-44f0-8b44-b148d8f75165. (period included in the link). Accessed Oct. 3, 2021.

Centers for Disease Control and Prevention (CDC). 2021b. COVID-19 Vaccines That Require 2 Shots. Aug. 9, 2021. Online at https://www.cdc.gov/coronavirus/2019-ncov/vaccines/second-shot.html. Accessed Oct. 3, 2021.

Centers for Disease Control and Prevention (CDC). 2021c. COVID-19 Vaccine Breakthrough Case Investigation and Reporting. Aug. 20, 2021. Online at https://www.cdc.gov/vaccines/covid-19/health-departments/breakthrough-cases.html. Accessed Oct. 3, 2021.

Centers for Disease Control and Prevention (CDC). 2021d. Selected Adverse Events Reported after COVID-19 Vaccination. Aug. 17, 2021. Online at https://www.cdc.gov/coronavirus/2019-ncov/vaccines/safety/adverse-events.html. Accessed Oct. 3, 2021.

Centers for Disease Control and Prevention (CDC). 2021e. COVID Data Tracker - Variant Proportions. Updated Sept. 22, 2021. Online at https://covid.cdc.gov/covid-data-tracker/#cases_totalcases. Accessed Oct. 3, 2021.

Callaway, E. 2021. Pfizer COVID vaccine protects against worrying coronavirus variants. Nature, May 6, 2021. Online at https://www.nature.com/articles/d41586-021-01222-5. Accessed Oct. 3, 2021.

European Medicines Agency (EMA). 2021. EMA recommends COVID-19 Vaccine AstraZeneca for authorization in the EU. Jan. 29, 2021. Online at https://www.ema.europa.eu/en/news/ema-recommends-covid-19-vaccine-astrazeneca-authorisation-eu. Accessed Oct. 3, 2021.

Evans, M. & J. Wernau. 2021. Unvaccinated Americans Are Behind Rising Covid-19 Hospitalizations. Wall Street Journal, July 18, 2021. Online at https://www.wsj.com/articles/unvaccinated-covid-19-hospitalizations-11626528110?mod=article_inline. Accessed Oct. 3, 2021.

Georgiou, A. 2021. How Contagious Are Chickenpox, Measles As CDC Document Reveals Delta Variant's R0. Newsweek, July 30, 2021. Online at https://www.newsweek.com/how-contagious-chickenpox-measles-cdc-document-delta-variant-coronavirus-r0-1614661. Accessed Oct. 3, 2021.Oct. 3

Griggs, M.B. 2014. 30,000 People In Quarantine After Bubonic Plague Kills One in China. Smithsonian, July 23, 2014. Online at https://www.smithsonianmag.com/smart-news/bubonic-plague-outbreak-china-leads-quarantine-180952136/. Accessed Oct. 3, 2021.

Haas, E.J., Angulo, F.J., McLaughlin, J.M., Anis, E., Singer, S.J., Khan, F., Brooks, N., Smaja, M. Mircus, G., Pan, K. Southern, J., Swerdlow, D.L., Jodar, L., Levy, Y., & Alroy-Preis, S. 2021. Impact and effectiveness of mRNA BNT162b2 vaccine against SARS-CoV-2 infections and COVID-19 cases, hospitalizations, and deaths following a nationwide vaccination campaign in Israel: an observational study using national surveillance data. The Lancet, May 05, 2021. Online at https://doi.org/10.1016/S0140-6736(21)00947-8. Accessed Oct. 3, 2021.

Hiltzik, M. 2021. Column: 'Death panels' arrive — in COVID-stricken Republican Idaho. Los Angeles Times, Sept. 17, 2021. Online at https://www.latimes.com/business/story/2021-09-17/death-panels-republican-covid-stricken-idaho. Accessed Oct. 3, 2021.

Johns Hopkins University & Medicine (JHUM). 2021. Coronavirus Resource Center: World Map >> US. Updated Oct. 3, 2021, 11:21 AM PT. Online at https://coronavirus.jhu.edu/map.html. Accessed Oct. 3, 2021.

Kates, J., Dawson, L., Anderson, A., Rouw, A., Michaud, J, & N. Singer. 2021. COVID-19 Vaccine Breakthrough Cases: Data from the States. Kaiser Family Foundation, July 30, 2021. Online at https://www.kff.org/policy-watch/covid-19-vaccine-breakthrough-cases-data-from-the-states/. Accessed Oct. 3, 2021.

Knowles, H. 2021. Hospitals overwhelmed by covid are turning to 'crisis standards of care.' What does that mean? Washington Post, Sept. 22, 2021. Online at https://www.washingtonpost.com/health/2021/09/22/crisis-standards-of-care/. Accessed Oct. 3, 2021.

Ledford, H., 2021. COVID vaccines and blood clots: five key questions. Nature, 592(7855), pp.495-496. Online at https://www.icpcovid.com/sites/default/files/2021-04/Ep%20132-12%20COVID%20vaccines%20and%20blood%20clots_%20five%20key%20questions.pdf. Accessed Oct. 3, 2021.

Liu, Y. and Rocklöv, J., 2021. The reproductive number of the Delta variant of SARS-CoV-2 is far higher compared to the ancestral SARS-CoV-2 virus. Journal of Travel Medicine. Online at https://academic.oup.com/jtm/advance-article/doi/10.1093/jtm/taab124/6346388. Accessed Oct. 3, 2021.

Lopez Bernal, J., Andrews, N., Gower, C., Gallagher, E., Simmons, R., Thelwall, S., Stowe, J., Tessier, E., Groves, N., Dabrera, G. and Myers, R., 2021. Effectiveness of Covid-19 vaccines against the B. 1.617. 2 (delta) variant. New England Journal of Medicine. Online at https://www.nejm.org/doi/full/10.1056/NEJMoa2108891. Accessed Oct. 3, 2021.

Luther, M. 2020. Whether One May Flee From A Deadly Plague. Aug. 1527. Reprinted in Christianity Today, May 19, 2020. Online at https://www.christianitytoday.com/ct/2020/may-web-only/martin-luther-plague-pandemic-coronavirus-covid-flee-letter.html. Accessed Oct. 3, 2021.

Mahase, E. 2020. Covid-19: Moderna vaccine is nearly 95% effective, trial involving high risk and elderly people shows. BMJ: British Medical Journal (Online), 371. Online at https://search.proquest.com/openview/f23612d9e0218b4d0ee67dcdb5c57884/1?pq-origsite=gscholar&cbl=2043523. Accessed Oct. 3, 2021.

Mallapaty, S. and Callaway, E., 2021. What scientists do and don't know about the Oxford-AstraZeneca COVID vaccine. Nature, 592(7852), pp.15-17. Online at https://www.icpcovid.com/sites/default/files/2021-03/Ep%20123-2%20What%20scientists%20do%20and%20don%E2%80%99t%20know%20about%20the%20Oxford%E2%80%93AstraZeneca%20COVID%20vaccine.pdf. Accessed Oct. 3, 2021.

National Weather Service (NWS). 2021. How Dangerous is Lightning? Online at https://www.weather.gov/safety/lightning-odds. Accessed Oct. 3, 2021.

Noor, R. 2021. Developmental Status of the Potential Vaccines for the Mitigation of the COVID-19 Pandemic and a Focus on the Effectiveness of the Pfizer-BioNTech and Moderna mRNA Vaccines. Current clinical microbiology reports, Mar. 3, 2021, pp.1-8. Online at https://link.springer.com/article/10.1007/s40588-021-00162-y. Accessed Oct. 3, 2021.

Oliver, S.E. et. al. 2020. The Advisory Committee on Immunization Practices' Interim Recommendation for Use of Moderna COVID-19 Vaccine—United States, December 2020. MMWR. Morbidity and mortality weekly report, 69. Online at https://www.cdc.gov/mmwr/volumes/69/wr/mm695152e1.htm. Accessed Oct. 3, 2021.

Olliaro, P., Torreele, E. and Vaillant, M. 2021. COVID-19 vaccine efficacy and effectiveness—the elephant (not) in the room. The Lancet Microbe. Online at https://www.thelancet.com/journals/lanmic/article/PIIS2666-5247(21)00069-0/fulltext?s=08. Accessed Oct. 3, 2021.

Polack, F.P., Thomas, S.J., Kitchin, N., Absalon, J., Gurtman, A., Lockhart, S., Perez, J.L., Pérez Marc, G., Moreira, E.D., Zerbini, C. and Bailey, R. 2020. Safety and efficacy of the BNT162b2 mRNA Covid-19 vaccine. New England Journal of Medicine, 383(27), pp.2603-2615. Online at https://www.nejm.org/doi/full/10.1056/NEJMoa2034577. Accessed Oct. 3, 2021.

Ritchie, H., Ortiz-Ospina, E., Beltekian, D., Mathieu, E., Hasell, J., Macdonald, B., Giattino, C., Appel, C., Rodés-Guirao, L., and M. Roser. 2021. Coronavirus (COVID-19) Vaccinations. Our World In Data, May 6, 2021. Online at https://ourworldindata.org/covid-vaccinations. Accessed Oct. 3, 2021.

Rosenberg ES, Holtgrave DR, Dorabawila V, et al. New COVID-19 Cases and Hospitalizations Among Adults, by Vaccination Status — New York, May 3–July 25, 2021. MMWR Morb Mortal Wkly Rep 2021;70:1306–1311. Online at http://dx.doi.org/10.15585/mmwr.mm7037a7. Accessed Oct. 3, 2021.

Rutkowski, K., Mirakian, R., Till, S., Rutkowski, R. and Wagner, A. 2021. Adverse reactions to COVID‐19 vaccines: a practical approach. Clinical & Experimental Allergy. Online at https://onlinelibrary.wiley.com/doi/full/10.1111/cea.13880. Accessed Oct. 3, 2021.

Schiferl, E., 1983. Iconography of Plague Saints in Fifteenth-century Italian Painting. Fifteenth Century Studies, 6, p.205. Online at https://www.proquest.com/openview/b63edc66b065c7024f2bf4ba28c1a661/1?pq-origsite=gscholar&cbl=1818258. Accessed Oct. 3, 2021.

Scobie HM, Johnson AG, Suthar AB, et al. Monitoring Incidence of COVID-19 Cases, Hospitalizations, and Deaths, by Vaccination Status — 13 U.S. Jurisdictions, April 4–July 17, 2021. MMWR Morb Mortal Wkly Rep 2021;70:1284–1290. Sept. 17, 2021. Online at https://www.cdc.gov/mmwr/volumes/70/wr/mm7037e1.htm?s_cid=mm7037e1_w. Accessed Oct. 3, 2021.

Self WH, Tenforde MW, Rhoads JP, et al. Comparative Effectiveness of Moderna, Pfizer-BioNTech, and Janssen (Johnson & Johnson) Vaccines in Preventing COVID-19 Hospitalizations Among Adults Without Immunocompromising Conditions — United States, March–August 2021. MMWR Morb Mortal Wkly Rep. ePub: 17 September 2021. Online at http://dx.doi.org/10.15585/mmwr.mm7038e1. Accessed Oct. 3, 2021.

Shimabukuro, T.T., Cole, M. and Su, J.R., 2021. Reports of anaphylaxis after receipt of mRNA COVID-19 vaccines in the US—December 14, 2020-January 18, 2021. Jama, 325(11), pp.1101-1102. Online at https://jamanetwork.com/journals/jama/article-abstract/2776557. Accessed Oct. 3, 2021.

Smout, A. 2021. English study finds 50-60% reduced risk of COVID for double-vaccinated. Reuters, Aug. 3, 2021. Online at https://www.reuters.com/world/uk/english-study-finds-50-60-reduced-risk-covid-double-vaccinated-2021-08-03/. Accessed Oct. 3, 2021.

Stamatatos, L., Czartoski, J., Wan, Y.H., Homad, L.J., Rubin, V., Glantz, H., Neradilek, M., Seydoux, E., Jennewein, M.F., MacCamy, A.J. and Feng, J., 2021. mRNA vaccination boosts cross-variant neutralizing antibodies elicited by SARS-CoV-2 infection. Science, Vol. 372, No. 6549. Online at https://science.sciencemag.org/content/372/6549/1413. Accessed Oct. 3, 2021.

Tanne, J.H., 2020. Covid-19: FDA panel votes to approve Pfizer BioNTech vaccine. BMJ 2020;371:m4799. Online at https://www.bmj.com/content/371/bmj.m4799. Accessed Oct. 3, 2021.

Tenforde, M.W. et. al. 2021. Effectiveness of Pfizer-BioNTech and Moderna Vaccines Against COVID-19 Among Hospitalized Adults Aged≥ 65 Years—United States, January–March 2021. MMWR. Morbidity and Mortality Weekly Report, 70. Online at https://www.cdc.gov/mmwr/volumes/70/wr/mm7018e1.htm. Accessed Oct. 3, 2021.

Tregoning, J.S., Flight, K.E., Higham, S.L. et al. Progress of the COVID-19 vaccine effort: viruses, vaccines and variants versus efficacy, effectiveness and escape. Nat Rev Immunol (2021). https://doi.org/10.1038/s41577-021-00592-1. Online at https://www.nature.com/articles/s41577-021-00592-1. Accessed Oct. 3, 2021.

University of New South Wales (UNSW). 2021. What We Now Know About the SARS-CoV-2 Delta Variant That's Wreaking Havoc Globally. Press release, Aug. 4, 2021. Online at https://scitechdaily.com/what-we-now-know-about-the-sars-cov-2-delta-variant-thats-wreaking-havoc-globally/. Accessed Oct. 3, 2021.

U.S. Food and Drug Administration (USFDA). 2020. Comirnaty and Pfizer-BioNTech COVID-19 Vaccine. Coronavirus Disease 2019 (COVID-19) press release, Aug. 23, 2021. Online at https://www.fda.gov/coronavirus-disease-2019-covid-19/comirnaty-and-pfizer-biontech-covid-19-vaccine. Accessed Oct. 3, 2021.

Westneat, D. 2021. 'Sophie’s choice, over and over': Death panels are the new phase of the pandemic. Seattle Times, Sept. 11, 2021. Online at https://www.seattletimes.com/seattle-news/health/sophies-choice-over-and-over-death-panels-are-the-new-phase-of-the-pandemic/. Accessed Oct. 3, 2021.

White, A. 2014. Plague and Pleasure: The Renaissance World of Pius II. CUA Press. ISBN: 0813226813, 9780813226811. Online at https://books.google.com/books?id=SfIdBgAAQBAJ. Accessed Oct. 3, 2021.

Posted in History, Theology, Uncategorized | 8 Comments