65 research outputs found

    A Closed System of Production Possibility and Social Welfare

    Get PDF
    We offer a closed system production possibility and social welfare system that can be modelled using virtually any available software package. It has the attribute that social welfare is not independent of production possibilities. The closure is made using the famous result by Negishi (1960) for a purely competitive economy. The goal is to help students to understand the interaction, through experimentation, between production and social choice in a competitive economy.

    Between Worlds: Securing Mixed JavaScript/ActionScript Multi-Party Web Content

    Get PDF
    Mixed Flash and JavaScript content has become increasingly prevalent; its purveyance of dynamic features unique to each platform has popularized it for myriad Web development projects. Although Flash and JavaScript security has been examined extensively, the security of untrusted content that combines both has received considerably less attention. This article considers this fusion in detail, outlining several practical scenarios that threaten the security of Web applications. The severity of these attacks warrants the development of new techniques that address the security of Flash-JavaScript content considered as a whole, in contrast to prior solutions that have examined Flash or JavaScript security individually. Toward this end, the article presents FlashJaX, a cross-platform solution that enforces fine-grained, history-based policies that span both Flash and JavaScript. Using in-lined reference monitoring, FlashJaX safely embeds untrusted JavaScript and Flash content in Web pages without modifying browser clients or using special plug-ins. The architecture of FlashJaX, its design and implementation, and a detailed security analysis are exposited. Experiments with advertisements from popular ad networks demonstrate that FlashJaX is transparent to policy-compliant advertisement content, yet blocks many common attack vectors that exploit the fusion of these Web platforms

    Automating Cyberdeception Evaluation with Deep Learning

    Get PDF
    A machine learning-based methodology is proposed and implemented for conducting evaluations of cyberdeceptive defenses with minimal human involvement. This avoids impediments associated with deceptive research on humans, maximizing the efficacy of automated evaluation before human subjects research must be undertaken. Leveraging recent advances in deep learning, the approach synthesizes realistic, interactive, and adaptive traffic for consumption by target web services. A case study applies the approach to evaluate an intrusion detection system equipped with application-layer embedded deceptive responses to attacks. Results demonstrate that synthesizing adaptive web traffic laced with evasive attacks powered by ensemble learning, online adaptive metric learning, and novel class detection to simulate skillful adversaries constitutes a challenging and aggressive test of cyberdeceptive defenses

    Experiences with Honey-Patching in Active Cyber Security Education

    Get PDF
    Abstract Modern cyber security educational programs that emphasize technical skills often omit or struggle to effectively teach the increasingly important science of cyber deception. A strategy for effectively communicating deceptive technical skills by leveraging the new paradigm of honeypatching is discussed and evaluated. Honey-patches mislead attackers into believing that failed attacks against software systems were successful. This facilitates a new form of penetration testing and capture-the-flag style exercise in which students must uncover and outwit the deception in order to successfully bypass the defense. Experiences creating and running the first educational lab to employ this new technique are discussed, and educational outcomes are examined

    Opaque control-flow integrity.

    Get PDF
    Abstract-A new binary software randomization and ControlFlow Integrity (CFI) enforcement system is presented, which is the first to efficiently resist code-reuse attacks launched by informed adversaries who possess full knowledge of the inmemory code layout of victim programs. The defense mitigates a recent wave of implementation disclosure attacks, by which adversaries can exfiltrate in-memory code details in order to prepare code-reuse attacks (e.g., Return-Oriented Programming (ROP) attacks) that bypass fine-grained randomization defenses. Such implementation-aware attacks defeat traditional fine-grained randomization by undermining its assumption that the randomized locations of abusable code gadgets remain secret. Opaque CFI (O-CFI) overcomes this weakness through a novel combination of fine-grained code-randomization and coarsegrained control-flow integrity checking. It conceals the graph of hijackable control-flow edges even from attackers who can view the complete stack, heap, and binary code of the victim process. For maximal efficiency, the integrity checks are implemented using instructions that will soon be hardware-accelerated on commodity x86-x64 processors. The approach is highly practical since it does not require a modified compiler and can protect legacy binaries without access to source code. Experiments using our fully functional prototype implementation show that O-CFI provides significant probabilistic protection against ROP attacks launched by adversaries with complete code layout knowledge, and exhibits only 4.7% mean performance overhead on current hardware (with further overhead reductions to follow on forthcoming Intel processors). I. MOTIVATION Code-reuse attacks (cf., Permission to freely reproduce all or part of this paper for noncommercial purposes is granted provided that copies bear this notice and the full citation on the first page. Reproduction for commercial purposes is strictly prohibited without the prior written consent of the Internet Society, the first-named author (for reproduction of an entire paper only), and the author's employer if the paper was prepared within the scope of employment. This has motivated copious work on defenses against codereuse threats. Prior defenses can generally be categorized into: CFI [1] and artificial software diversity CFI restricts all of a program's runtime control-flows to a graph of whitelisted control-flow edges. Usually the graph is derived from the semantics of the program source code or a conservative disassembly of its binary code. As a result, CFIprotected programs reject control-flow hijacks that attempt to traverse edges not supported by the original program's semantics. Fine-grained CFI monitors indirect control-flows precisely; for example, function callees must return to their exact callers. Although such precision provides the highest security, it also tends to incur high performance overheads (e.g., 21% for precise caller-callee return-matching [1]). Because this overhead is often too high for industry adoption, researchers have proposed many optimized, coarser-grained variants of CFI. Coarse-grained CFI trades some security for better performance by reducing the precision of the checks. For example, functions must return to valid call sites (but not necessarily to the particular site that invoked the callee). Unfortunately, such relaxations have proved dangerous-a number of recent proof-of-concept exploits have shown how even minor relaxations of the control-flow policy can be exploited to effect attacks Artificial software diversity offers a different but complementary approach that randomizes programs in such a way that attacks succeeding against one program instance have a very low probability of success against other (independently randomized) instances of the same program. Probabilistic defenses rely on memory secrecy-i.e., the effects of randomization must remain hidden from attackers. One of the simplest and most widely adopted forms of artificial diversity is Address Space Layout Randomization (ASLR), which randomizes the base addresses of program segments at loadtime. Unfortunately, merely randomizing the base addresses does not yield sufficient entropy to preserve memory secrecy in many cases; there are numerous successful derandomization attacks against ASLR Recently, a new wave of implementation disclosure attacks Experiments show that O-CFI enjoys performance overheads comparable to standard fine-grained diversity and non-opaque, coarse-grained CFI. Moreover, O-CFI's control-flow checking logic is implemented using Intel x86/x64 memory-protection extensions (MPX) that are expected to be hardware-accelerated in commodity CPUs from 2015 onwards. We therefore expect even better performance for O-CFI in the near future. Our contributions are as follows: • We introduce O-CFI, the first low-overhead code-reuse defense that tolerates implementation disclosures. • We describe our implementation of a fully functional prototype that protects stripped, x86 legacy binaries without source code. II. THREAT MODEL Our work is motivated by the emergence of attacks against fine-grained diversity and coarse-grained control-flow integrity. We therefore introduce these attacks and distill them into a single, unified threat model. A. Bypassing Coarse-Grained CFI Ideally, CFI permits only programmer-intended control-flow transfers during a program's execution. The typical approach is to assign a unique ID to each permissible indirect controlflow target, and check the IDs at runtime. Unfortunately, this introduces performance overhead proportional to the degree of the graph-the more overlaps between valid target sets of indirect branch instructions, the more IDs must be stored and checked at each branch. Moreover, perfect CFI cannot be realized with a purely static control-flow graph; for example, the permissible destinations of function returns depend on the calling context, which is only known at runtime. Fine-grained CFI therefore implements a dynamically computed shadow stack, incurring high overheads To avoid this, coarse-grained CFI implementations resort to a reduced-degree, static approximation of the control-flow graph, and merge identifiers at the cost of reduced security. Typically, attackers need more than a single 4K page worth of code to find enough gadgets to mount a code-reuse attack. To discourage brute-force searches for more code pages, artificial diversity defenses routinely mine the address space with unmapped pages that abort the process if accessed B. Assumptions Given these sobering realities, we adopt a conservative threat model that assumes that attackers will eventually find and disassemble all code pages in victim processes. Our threat model therefore assumes that the adversary knows the complete in-memory code layout-including the locations of any gadgets required to launch a ROP attack. We also assume that the attacker can read and write the full contents of the heap and stack, as well as any data structures used by the dynamic loader. In keeping with common practice, we assume that data execution protection is activated, so that code page permissions can be maintained as either writable or executable but not both. However, we assume that attackers cannot safely perform a comprehensive, linear scan of virtual memory, since defenders may place unmapped guard pages at random locations. Instead, attackers must follow references from one disclosed memory page to another III. O-CFI OVERVIEW O-CFI combines insights from CFI and automated software diversity. It extends CFI with a new, coarse-grained CFI enforcement strategy inspired by bounds-checking, that validates control-flow transfers without divulging the bounds against which their destinations are checked. Bounds-checking is fast, the bounds are easier to conceal than arbitrary gadget locations, and the bounds are randomizable. This imbues CFI and fine-grained software diversity with an additional layer of protection against code-reuse attacks aided by implementation disclosures. As a result, O-CFI enjoys performance similar to coarse-grained CFI, with probabilistic security guarantees similar to fine-grained artificial diversity in the absence of implementation disclosures. Following traditional CFI, an O-CFI policy assigns to each indirect branch site a destination set that captures its set of permissible destination addresses. Such a graph can be derived from the program's source code or (with lesser precision) a conservative disassembly of its object code. We next reformulate this policy as a bounds-checking problem by reducing each destination set to only its minimal and maximal members. This policy approximation can be efficiently enforced by confining each branch to the memory-aligned addresses within its destination set range. All intended destination addresses are aligned within these bounds, so the enforcement conservatively preserves intended control-flows. Code layout is optimized to tighten the bounds, so that the set of unintended, aligned destinations within the bounds remains minimal. These few remaining unintended but reachable destinations are protected by the artificial diversity half of our approach. Our artificial diversity approach probabilistically protects the aligned, in-bounds, but policy-violating control-flows by applying fine-grained randomization to the binary code at load-time. While the overall strategy for implementing this randomization step is based on prior works Reformulating CFI in this way forces attackers to change their plan of attack. The recent attacks against coarse-grained CFI succeed by finding exploitable code that is reachable due to policy-relaxations needed for acceptable performance. These relaxations admit an alarming array of false-positives: instead of identifying the actual caller, all call-preceded instructions are incorrectly identified as permitted branch destinations. Such instructions saturate a typical address space, giving attackers too much wiggle room to build attacks. O-CFI counters this by changing the approximation approach: each branch destination is restricted to a relatively short span of aligned addresses, with all the bounds chosen pseudo-randomly at load-time. This greatly narrows the field of possible hijacks, and it removes the opportunity for attackers to analyze programs ahead of time for viable ROP gadget chains. In O-CFI, no two program instances admit the same set of ROP payloads, since the bounds are all randomized every time the program is loaded. Since the security of coarse-grained CFI depends in part on the precision of its policy approximation, it is worthwhile to improve the precision by tightening the bounds imposed upon each branch. This effectively reduces the space of attacker guesses that might succeed in hijacking any given branch. To reduce this space as much as possible, we introduce a novel binary code optimization, called portals, that minimizes the distance covered by the lowest and greatest element of each indirect branch's destination set. Our fine-grained artificial diversity implementation is an adaptation and extension of binary stirring To protect against information leaks that might disclose bounds information, our implementation is carefully designed to keep all bounds opaque to external threats. They are randomly chosen at load-time (as a side-effect of binary stirring) and stored in a bounds lookup table (BLT) located at a randomly chosen base address. The table size is very small relative to the virtual address space, and attackers cannot safely perform bruteforce scans of the full address space (see §II-B), so guessing the BLT's location is probabilistically infeasible for attackers. No code or data sections contain any pointer references to BLT addresses; all references are computed dynamically at load-time and stored henceforth exclusively in protected registers. A. Bounding the Control Flow For each indirect branch site with (non-empty) destination set D, O-CFI guards the branch instruction with a bounds-check that continues execution only if the impending target t satisfies t ∈ [min D, max D]. Indirect branch instructions include all control-flow transfer instructions that target computed destinations, including return instructions. Failure of the boundscheck solicits immediate process termination with an error code (for easier debugging). Termination could be replaced with a different intervention if desired, such as an automated attack analysis or alarm, followed by restart and re-randomization. The bounds-check implementation first loads the pair (min D, max D) from the BLT into registers via an indirect, indexed memory reference. The load instruction's arguments and syntax are independent of the BLT's location, concealing its address from attackers who can read the checking code. The impending branch target t is then checked against the loaded bounds. If the check succeeds, execution continues; otherwise the process immediately terminates with a bounds range (#BR) exception. The #BR exception helps distinguish between crashes and guessing attacks. To resist guessing attacks (e.g., BROP), web servers and other services should use this exception to trigger re-randomization as they restart. Following the approaches of PittSFIeld To bypass these checks, an attacker must craft a payload whose every gadget is properly aligned and falls within the bounds of the preceding gadget's conclusory indirect branch. The odds of guessing a reachable series of such gadgets decrease exponentially with the number of gadgets in the desired payload. B. Opacifying Control-flow Bounds Diversifying bounds. The bounds introduced by O-CFI constitute a coarse-grained CFI policy. Section II warns that such coarse granularity can lead to vulnerabilities. However, to exploit such vulnerabilities, attackers must discover which control-flows adhere to the CFI policy and which do not. To make the impermissible flows opaque to attackers, we use diversity. Our prototype uses a modified version of the technique outlined by Wartell et al. Performing fine-grain code randomization at load-time indirectly randomizes the ranges used to bound the control-flow. In contrast to other CFI techniques, attackers therefore do not have a priori knowledge of the control-flow bounds. Preventing Information Leaks. Attackers bypass fine grained diversity using information leaks, such as those described in §II-A. Were O-CFI's control-flow bounds expressed as constants in the instruction stream, attackers could bypass our defense via information leaks. To avoid this, we instead confine this sensitive information to an isolated data page, the BLT. The BLT is initialized at a random virtual memory address at load-time, and there are no pointer references (obfuscated or otherwise) to any BLT address in any code or data page in the process. This keeps its location hidden from attackers. Furthermore, we take additional steps to prevent accidental BLT disclosure via pointer leaks. Our prototype stores BLT base addresses in segment selectors-a legacy feature of all x86 processors. In particular, each load from the BLT uses the gs segment selector and a unique index to read the correct bounds. We only use the gs selector for instructions that implement bounds checks, so there are no other instructions that adversaries can reuse to learn its value. Attackers are also prevented from executing instructions that reveal the contents of the segment registers, since such instructions are privileged. To succeed, attackers must therefore (i) guess branch ranges, or (ii) guess the base address of the BLT. The odds of correctly guessing the location of the BLT are low enough to provide probabilistic protection. On 32-bit Windows Systems, for instance, the chances of guessing the base address are or less than one in two billion. Incorrect guesses alert defenders and trigger re-randomization with high probability (by accessing an unallocated memory page). The likelihood of successfully guessing a reachable gadget chain is a function of the length of the chain and the span of the bounds. The next section therefore focuses on reducing the average bounds span. C. Tightening Control-flow Check Bounds The distance between the lowest and highest intended destinations of any given indirect control-flow transfer instruction depends on the code layout. Placing indirect branches close to their targets both reduces bounds and improves locality, elevating both security and efficiency. Therefore we organize the code segment into clusters-one per indirect branch-each containing the basic blocks targeted by a particular branch. To accommodate blocks that are destinations of multiple distinct branch instructions, we consider three options: (i) put the block in one cluster and expand the bounds of other branches to include its address, (ii) create duplicate copies of the block in multiple clusters, or (iii) add a portal block to each cluster, which unconditionally jumps to the block. Each solution incurs a trade-off: expanding bounds reduces security, creating duplicates increases code size, and portals introduce runtime overhead. The options are not mutually exclusive, affording optimizers a range of strategies. Our experiments indicate that portals are often the best choice (see below). The capacity of the portal system limits the number of portals per nexus. Varying nexus capacity allows O-CFI to be tuned to different requirements. Setting it to zero prevents the creation of any portals, forcing the optimizer to choose alternative options. At the other extreme, setting no upper limit allows a portal to be created for every target, reducing all bounds ranges to wt, where w is the alignment width (usually 16 bytes; see §V-A) and t is the number of targets of the branch. At this setting, all indirect branches can only branch into a nexus, and through them, only to exactly those addresses that have been statically identified as targets. Thus, O-CFI with unbounded nexus capacity enforces fine-grained, static CFI. The extra layer of indirection imposed by a portal has a minor impact on runtime; there is thus a trade-off between security and performance. Users may opt for full CFI enforcement with O-CFI for security-critical components, and lower the nexus capacity to a desired performance level for less critical software. In our experiments, we found that a nexus capacity of 12 results in a significant reduction in bounds sizes with imperceptible performance effects. All of our experiments in §V use this nexus capacity. Section V-D details how different nexus capacities affect bound ranges. D. Example Defense against JIT-ROP The following example illustrates how O-CFI secures binaries against disclosure attacks. Consider a binary whose code segment contains five useful gadgets g 1 , . . . , g 5 . Each gadget terminates in an indirect branch protected by a bounds check. Under appropriate conditions, a disclosure attack such as JIT-ROP is able to recover a large portion of the runtime layout of the binary In our example, if g 1 is selected to be part of the payload, it can only be chained with gadget g 4 or g 5 . Attempting to jump from g 1 to any other gadget triggers a bounds violation that stops the attack. Similarly, an attack that hijacks a controlflow to c 1 can only redirect it to gadgets g 1 , g 2 , or g 3 ; all other gadgets are outside cluster c 1 and are therefore detected as impermissible destinations of the hijacked branch. Broadly speaking, all links in a payload's chain must traverse edges in the Cartesian product of the (aligned) gadget sets within the corresponding clusters. A successful attack must therefore limit itself to an extremely sparse graph of available edges. Our experiments (see §V-C) indicate that in practice the probability of successfully chaining gadgets in such a sparse graph is very low-just 0.01% for a four-gadget payload. The entropy of our procedure is further analyzed in §VI-A. IV. O-CFI IMPLEMENTATION We have implemented a fully functional prototype of O-CFI for the Intel x86 architecture. Our implementation uses a binary rewriting framework that secures COTS x86 binaries without source, debug, or relocation information. Like traditional CFI, however, we emphasize that O-CFI is equally suitable for inclusion in a compiler. Our rewriter generates a transformed version of the binary that leverages 1) a coarse-grained CFI policy that bounds control-flows, 2) fine-grained randomization to thwart traditional ROP attacks and diversify control-flow bounds so they become unknown and unreliable for attackers, 3) x86 segmentation registers to prevent accidental leakage of the bounds lookup table (BLT), and 4) an SFI framework similar to PittSFIeld [29] to enforce instruction alignment, denying attackers access to misaligned instructions that bypass bounds checks. The architecture of O-CFI is shown in A. Static Binary Rewriting 1) Conservative Disassembly: We first disassemble the code section using a conservative disassembler. Similar to the approach outlined by Wartell et al. [45], the code section is duplicated, with the old copy (renamed to .told) serving as a read-only data segment and the new copy (called .tnew) containing the rewritten executable code. The .told section is set non-executable, and all code blocks identified as possible targets of indirect jumps are overwritten with a five-byte tagged pointer. The tagged pointer consists of a tag byte (0xF4) followed by the four-byte address of that block in the .tnew section. The tag byte facilitates efficient runtime redirection of stale pointers to their correct targets,
    corecore